mechanical, thermal and structural investigations of one

231
Mechanical, Thermal and Structural Investigations of One-Part Geopolymer Concrete A thesis by Mahya Askarian Supervisors: Professor Bijan Samali Professor Zhong Tao Professor Georgius Adam A thesis submitted in fulfilment of the requirements for the degree of Doctor of Philosophy Centre for Infrastructure Engineering, Western Sydney University 2019

Upload: others

Post on 16-Apr-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Mechanical, Thermal and Structural Investigations of One

Mechanical, Thermal and Structural

Investigations of One-Part Geopolymer

Concrete

A thesis by

Mahya Askarian

Supervisors:

Professor Bijan Samali

Professor Zhong Tao

Professor Georgius Adam

A thesis submitted in fulfilment of the requirements for the degree of

Doctor of Philosophy

Centre for Infrastructure Engineering,

Western Sydney University

2019

Page 2: Mechanical, Thermal and Structural Investigations of One
Page 3: Mechanical, Thermal and Structural Investigations of One

ii

Abstract

Geopolymeric materials have recently attracted increasing research due to the need to

reduce global CO2 emission as well as the growing desire for utilisation of waste and

by-products in construction materials. Furthermore, geopolymers are well-known for

their excellent behaviour in terms of fire performance which is a matter of great

concern in structural design. This thesis presents a study on the development of one-

part geopolymer concrete (OGC) and hybrid OPC-geopolymer concrete (HGC)

mixes and their behaviour at both ambient and elevated temperatures and their

application in reinforced concrete columns. The intention for development of one-

part geopolymer concrete is to improve the commercial viability of geopolymers in

building industry. The previous studies mainly emphasized on the investigation of

geopolymeric material properties at ambient and elevated temperatures and only a

few experimental studies have been devoted to explore the fire performance of

reinforced geopolymer concrete members. Thus, this research aims to address this

research gaps.

In this study, material properties such as slump, density, compressive strength,

modulus of elasticity, tensile strength and stress-strain behaviour of OGC and HGC

mixes are compared with those of reference Ordinary Portland Cement concrete (CC)

mix. Furthermore, the hot compressive strength and thermal properties of the

mentioned mixes were measured.

Microstructural characterisation was also performed on OGC, HGC and CC samples

using scanning electron microscopy (SEM), energy dispersive X-ray spectroscopy

(EDS), X-ray diffraction (XRD) and Fourier-transform infrared spectroscopy (FTIR)

to provide better insight into microstructure of materials with distinct properties.

Page 4: Mechanical, Thermal and Structural Investigations of One

iii

Two kinds of structural tests were performed on the reinforced geopolymer concrete

columns. Two reinforced columns made of OGC and HGC were loaded axially at

ambient temperature. The effects of concrete type on the failure modes, development

of axial and lateral strains and ultimate load capacity of reinforced columns are

examined. In addition, the test results are compared with finite element (FE) analysis

for HGC and OGC compared with CC columns. Another ten columns including 3

OGC, 3 HGC and 4 CC columns were tested in fire in which each column was

exposed to a given constant load level at ambient temperature, then heated until the

column failed. The effects of concrete type (OGC, HGC and CC) and load level (0.3-

0.45 of ultimate) were studied in the test program. The experimental results

demonstrated that the use of geopolymer concrete has the potential to enhance the

fire resistance of reinforced columns. Furthermore, the achieved data from the

experiments such as temperature development, axial and lateral deformation, fire

resistance time and failure modes have improved the understanding of the

performance of reinforced OGC and HGC columns subjected to fire condition.

The test data was also used to verify the proposed finite element (FE) modelling

approach in which a finite element model, using ABAQUS commercial software,

simulates the behaviour of reinforced OGC, HGC and CC columns. A reasonable

agreement was observed by comparing the results obtained from experiments and FE

analyses.

This research confirms the practicality of employing sustainable one-part

geopolymer concrete in structural members and its benefits to improve the thermal

insulation or fire performance of reinforced concrete members.

Page 5: Mechanical, Thermal and Structural Investigations of One

iv

Acknowledgements

I would like to express my deepest gratitude and appreciation to the individuals who

supported me during my PhD studies. This research work could have not been

possible without their support and guidance.

First of all, I would like to thank my principal supervisor, Professor Bijan Samali, for

his dedicated supervision, invaluable guidance and limitless support during the

period of candidature. I have been greatly encouraged by his endless pursuit of

excellence and hard work. I will be grateful to him forever and I am very glad of this

honour to be his student.

I greatly appreciate my co-supervisors, Professor Zhong Tao and Professor Georgius

Adam for their mentorships, valuable recommendations, encouragement and

continuous support throughout the course of this research

I would like to thank all the staff in Centre for Infrastructure Engineering, Western

Sydney University. I greatly acknowledge the kind assistance from the technical staff

in the structure laboratory including Mr. Robert Marshall, Mr. Zachariah White, Mr.

Murray Bolden and Mr. Ali Gharizadeh. I also would like to express my special

thanks to Dr. Richard Wuhrer and Dr. Shamila Salek for their assistance in using the

advanced materials characterization facility (AMCF) at WSU. I am also appreciative

to Dr. Zhu Pan for his valuable support and assistance during this research.

I would like to take the opportunity to express my heartfelt gratitude to my kind and

supportive parents and my lovely sisters for their priceless support, continuous love

and confidence in me. Special gratitude and love to my husband for his continuous

patience and support and for being by my side, and always giving me reasons to

cheer. Without the moral and emotional support of my family, this work would not

have been possible.

Page 6: Mechanical, Thermal and Structural Investigations of One

v

Preface

This thesis has been supported by papers that have been published in internationally

recognised journals and conferences. These papers are listed below:

Journal Papers

• M Askarian, Z Tao, G Adam and B Samali. Mechanical properties of ambient

cured one-part hybrid OPC-geopolymer concrete. Construction and Building

Materials, 2018. 186: p. 330-337.

• M Askarian, Z Tao, B Samali and G Adam. Mix composition and

Characterisation of one-part geopolymers with different activators.

Construction and Building Materials, 2019. 225: p. 526-537.

• M Askarian, Z Tao, B Samali and G Adam. Fire performance of one-part

geopolymer columns (under preparation)

Conference Papers

• M Askarian, B Samali, Z Tao and G Adam. Mechanical and thermal

properties of hybrid OPC-geopolymer concrete. Concrete 2019, Sydney,

Australia.

• M Askarian, B Samali, Z Tao and G Adam. Microstructural Characterisation

of one-part geopolymers. Concrete 2019, Sydney, Australia.

Page 7: Mechanical, Thermal and Structural Investigations of One

vi

Contents

Statement of Authentication ..................................................................................................... i

Abstract .................................................................................................................................. ii

Acknowledgements ................................................................................................................ iv

Preface .................................................................................................................................... v

Contents ................................................................................................................................. vi

List of Tables .......................................................................................................................... x

List of Figures ........................................................................................................................ xi

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

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

1.2 Overview and Significance ........................................................................................... 1

1.3 Research Objectives and Scope of Work ....................................................................... 5

1.4 Thesis Layout ................................................................................................................ 7

Chapter 2 Literature Review ................................................................................................. 10

2.1 General ........................................................................................................................ 10

2.2 Source Materials ......................................................................................................... 10

2.3 Alkali Activators ......................................................................................................... 13

2.4 Curing Conditions ....................................................................................................... 14

2.5 Behaviour of Geopolymer Concrete at Ambient Temperature .................................... 15

2.5.1 Fresh properties .................................................................................................... 15

2.5.2 Mechanical properties .......................................................................................... 16

2.5.3 Microstructure of geopolymers ............................................................................ 17

2.6 Properties of Geopolymers at Elevated Temperatures ................................................. 20

2.6.1 Spalling ................................................................................................................ 20

2.6.2 Mechanical properties .......................................................................................... 21

2.6.3 Thermal properties ............................................................................................... 23

2.7 Recent Developments in Geopolymers ....................................................................... 24

2.7.1 Blended geopolymers ........................................................................................... 24

2.7.2 One-part geopolymers .......................................................................................... 26

2.8 Performance of Geopolymer Concrete Members at Ambient Temperature ................. 29

2.8.1 Geopolymer concrete beams ................................................................................ 29

2.8.2 Geopolymer concrete columns ............................................................................. 35

2.8.3 Other geopolymer concrete members ................................................................... 44

2.9 Fire Resistance of Reinforced Concrete Members ...................................................... 46

Page 8: Mechanical, Thermal and Structural Investigations of One

vii

2.9.1 Philosophy ............................................................................................................ 46

2.9.2 Spalling ................................................................................................................ 47

2.9.3 Fire resistance tests on reinforced OPC concrete columns ................................... 48

2.9.4 Fire resistance of geopolymer concrete elements ................................................. 52

2.10 Concluding Remarks ................................................................................................. 55

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete ........................................... 57

3.1 General ........................................................................................................................ 57

3.2 Experimental Procedure .............................................................................................. 57

3.2.1 Materials............................................................................................................... 57

3.2.2 Concrete mix proportions and specimen preparation ............................................ 58

3.2.3 Testing procedure ................................................................................................. 62

3.3 Results and Discussion ................................................................................................ 64

3.3.1 Behaviour of fresh mixes...................................................................................... 64

3.3.2 Compressive strength ........................................................................................... 66

3.3.3 Microstructural analysis ....................................................................................... 70

3.4 Conclusions ................................................................................................................. 74

Chapter 4 Development of One-Part Geopolymers ............................................................... 77

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

4.2 Theoretical Background .............................................................................................. 77

4.3 Experimental Procedure .............................................................................................. 79

4.3.1 Materials............................................................................................................... 79

4.3.2 Mix proportions and mixing ................................................................................. 80

4.3.3 Testing procedure ................................................................................................. 84

4.4 Results and Discussion ................................................................................................ 85

4.4.1 Workability .......................................................................................................... 85

4.4.2 Compressive strength and density ........................................................................ 86

4.4.3 FTIR Results ........................................................................................................ 91

4.4.4 XRD patterns ........................................................................................................ 93

4.4.5 SEM/EDS results.................................................................................................. 96

4.5 Conclusions ................................................................................................................. 99

Chapter 5 Mechanical and Thermal Properties ................................................................... 102

5.1 General ...................................................................................................................... 102

5.2 Experimental Procedure ............................................................................................ 102

5.2.1 Materials............................................................................................................. 102

5.2.2 Mix proportions and mixing ............................................................................... 104

5.2.3 Testing procedure ............................................................................................... 106

Page 9: Mechanical, Thermal and Structural Investigations of One

viii

5.3 Test Results and Discussion ...................................................................................... 114

5.3.1 Mechanical properties at ambient temperature ................................................... 114

5.3.2 Mechanical properties at elevated temperatures ................................................. 117

5.3.3 Thermal properties ............................................................................................. 122

5.4 Evaluation of Test Results with Code Predictions ..................................................... 125

5.4.1 Modulus of elasticity (MOE) .............................................................................. 125

5.4.2 Stress-strain relationship of concrete .................................................................. 127

5.4.3 Compressive reduction factors at elevated temperatures .................................... 129

5.4.4 Thermal conductivity ......................................................................................... 130

5.5 Conclusions ............................................................................................................... 130

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature ........... 134

6.1 General ...................................................................................................................... 134

6.2 Experimental Program............................................................................................... 135

6.2.1 Specimen preparation and material properties .................................................... 135

6.2.2 Test procedure and setup .................................................................................... 138

6.3 Experimental Results and Discussion ........................................................................ 141

6.3.1 General observations and failure modes ............................................................. 141

6.3.2 Axial load versus axial displacement curves ...................................................... 142

6.3.3 Lateral deformation ............................................................................................ 143

6.4 Evaluation of Test Results with FE Predictions ........................................................ 144

6.4.1 Materials modelling ............................................................................................ 144

6.4.2 Other details of FE model................................................................................... 146

6.4.3 Comparison of predictions with experimental results ......................................... 147

6.4.4 Sensitivity analysis ............................................................................................. 150

6.4.5 Verification of current test results ...................................................................... 151

6.5 Conclusions ............................................................................................................... 154

Chapter 7 Performance of Reinforced Concrete Columns at Elevated Temperatures ......... 156

7.1 General ...................................................................................................................... 156

7.2 Experimental Program............................................................................................... 157

7.2.1 Specimen preparation and material properties .................................................... 157

7.2.2 Test procedure and set up ................................................................................... 159

7.3 Test Results and Discussion ...................................................................................... 161

7.3.1 General observations and failure modes ............................................................. 161

7.3.2 Temperature versus time curves ......................................................................... 171

7.3.3 Axial deformation .............................................................................................. 174

7.3.4 Fire resistance..................................................................................................... 176

Page 10: Mechanical, Thermal and Structural Investigations of One

ix

7.4 Evaluation of Test Results with FE Predictions ........................................................ 178

7.4.1 Thermal analysis ................................................................................................ 179

7.4.2 Stress analysis .................................................................................................... 180

7.4.3 Comparison of predictions with experimental results ......................................... 185

7.5 Conclusions ............................................................................................................... 191

Chapter 8 Conclusions and Future Research Needs ............................................................ 194

8.1 Conclusions ............................................................................................................... 194

8.2 Recommendations for Future Works ......................................................................... 197

References ........................................................................................................................... 200

Page 11: Mechanical, Thermal and Structural Investigations of One

x

List of Tables

Table 2.1 Beam details ......................................................................................................... 31

Table 2.2 Column details ...................................................................................................... 37

Table 2.3 Test variables of the columns .............................................................................. 41

Table 3.1 Chemical composition of OPC, fly ash and slag determined by XRF................... 58

Table 3.2 Mix proportions of one-part geopolymer concrete ............................................... 61

Table 3.3 Compressive strengths of geopolymer concrete mixes at 1, 7 and 28 days ........... 68

Table 3.4 Oxide molar ratios in different mixes ................................................................... 72

Table 4.1 Chemical composition of OPC, fly ash and slag determined by XRF................... 79

Table 4.2 Mix proportions of one-part geopolymer concrete ............................................... 83

Table 4.3 Test results of geopolymer concrete mixes. .......................................................... 88

Table 5.1 Chemical composition of OPC, fly ash and slag determined by XRF................. 103

Table 5.2 Details of mix proportions ................................................................................. 104

Table 5.3 Mechanical properties of concrete mixes ............................................................ 115

Table 6.1 Material properties of reinforced steel ................................................................ 137

Table 6.2 Details of reinforced columns ............................................................................. 140

Table 6.3 Summary of the experimental test results and numerical modelling ................... 149

Table 6.4 Summary of sensitivity analysis parameters and results ..................................... 151

Table 6.5 Summary of the experimental results by Sumajouw et al. and numerical modelling

of OGCI-1 and OGCI-4 specimens ..................................................................................... 153

Table 7.1 Details of concrete columns tested in fire ........................................................... 159

Table 7.2 Measured Fire resistance of the columns ............................................................ 178

Table 7.3 Eurocode 3 parameters and formulations for carbon steel properties.................. 183

Table 7.4 Comparison of fire resistance of columns ........................................................... 191

Page 12: Mechanical, Thermal and Structural Investigations of One

xi

List of Figures

Figure 1.1 Types of polysialate .............................................................................................. 2

Figure 2.1 SEM micrograph of Class F fly ash particles ...................................................... 11

Figure 2.2 SEM micrograph of GGBFS ............................................................................... 11

Figure 2.3 SEM micrograph of metakaolin .......................................................................... 12

Figure 2.4 XRD pattern of fly ash based geopolymer .......................................................... 18

Figure 2.5 IR spectrum of fly ash and fly ash based geopolymer ......................................... 19

Figure 2.6 XRD pattern of geopolymers with varying slag content...................................... 20

Figure 2.7 Geopolymer concrete beams after demoulding ................................................... 30

Figure 2.8 Test sample detail of beams ................................................................................ 32

Figure 2.9 Reinforcement of beam in ABAQUS .................................................................. 34

Figure 2.10 Column details and set-up ................................................................................. 36

Figure 2.11 Column in the test machine ............................................................................... 36

Figure 2.12 Failure mode of geopolymer concrete columns ................................................. 38

Figure 2.13 Load-deflection curves for different tested geopolymer concrete columns ....... 39

Figure 2.14 Configuration of circular geopolymer concrete columns .................................. 40

Figure 2.15 column test set-up ............................................................................................. 42

Figure 2.16 Details and configuration of the column specimens .......................................... 43

Figure 2.17 Configuration of the columns ............................................................................ 43

Figure 2.18 Failure modes of columns ................................................................................. 44

Figure 2.19 Reinforcement arrangement in the panel ........................................................... 45

Figure 2.20 Elevation and cross section of reinforced concrete columns ............................ 50

Figure 2.21 Elevation and cross-section of the specimens.................................................... 52

Figure 2.22 A concrete test panel set for fire exposure......................................................... 53

Figure 2.23 Failure of the 150 mm OPC concrete panel....................................................... 54

Figure 2.24 Failure of the 150 mm geopolymer concrete panel ........................................... 54

Figure 3.1 Mixing one-part hybrid OPC-geopolymer concrete ............................................ 60

Figure 3.2 Concrete cylinders cast in steel cylindrical moulds ............................................. 61

Figure 3.3 Conducting slump test for one-part hybrid OPC-geopolymer concrete mixes .... 62

Figure 3.4 Vicat apparatus for measuring setting time of paste ............................................ 63

Figure 3.5 Test setup for compressive strength test .............................................................. 63

Page 13: Mechanical, Thermal and Structural Investigations of One

xii

Figure 3.6 Effect of OPC inclusion on the slump ................................................................. 65

Figure 3.7 Setting times of one-part pastes........................................................................... 66

Figure 3.8 Influence of calcium hydroxide and sodium silicate on strength development.... 69

Figure 3.9 BSE images of hybrid OPC-geopolymer mixes .................................................. 71

Figure 3.10 XRD analysis of various hybrid OPC-geopolymer systems at the age of 28 days.

.............................................................................................................................................. 74

Figure 4.1 Sample preparation ............................................................................................. 83

Figure 4.2 Relative slump values of one-part geopolymer mixes ......................................... 86

Figure 4.3 FTIR spectra of raw materials and typical geopolymer mixes ............................ 93

Figure 4.4 XRD patterns of raw materials and typical geopolymer mixes ........................... 95

Figure 4.5 BSE images of fly ash based one-part geopolymer mixes. .................................. 98

Figure 4.6 BSE images of fly ash/slag based one-part geopolymer mixes. .......................... 98

Figure 4.7 SEM images of geopolymer mixes of M1 and M11. ........................................... 99

Figure 5.1 Mixing concrete in laboratory ........................................................................... 105

Figure 5.2 Demoulded specimens after 24 h ...................................................................... 106

Figure 5.3 MOE test setup .................................................................................................. 108

Figure 5.4 Splitting tensile test ........................................................................................... 109

Figure 5.5 Test setup for MOR test .................................................................................... 110

Figure 5.6 Test setup for compressive stress-strain test ...................................................... 111

Figure 5.7 Test setup for investigation of compressive strength of samples at elevated

temperatures ........................................................................................................................ 112

Figure 5.8 The thermal properties measurement according to TPS .................................... 114

Figure 5.9 Stress-strain behaviour of concrete mixes after 28 days .................................... 116

Figure 5.10 Comparison of normalised stress-strain curves for different concrete mixes .. 117

Figure 5.11 Comparison of the reduction factors of compressive strength for OGC, HGC

and CC at various elevated temperatures ............................................................................ 119

Figure 5.12 Strain at peak stress for OGC, HGC and CC at various elevated Temperatures

............................................................................................................................................ 120

Figure 5.13 Effect of elevated temperatures on the physical appearance of one-part

geopolymer concrete (OGC) ............................................................................................... 121

Figure 5.14 Effect of elevated temperatures on the physical appearance of one-part hybrid

OPC-geopolymer concrete (HGC) ...................................................................................... 121

Figure 5.15 Effect of elevated temperatures on the physical appearance of OPC concrete

(CC) .................................................................................................................................... 122

Page 14: Mechanical, Thermal and Structural Investigations of One

xiii

Figure 5.16 Effect of temperature on thermal conductivity ................................................ 123

Figure 5.17 Effect of temperature on specific heat ............................................................. 125

Figure 5.18 Measured MOE against predictions ................................................................ 126

Figure 5.19 Stress-strain curves of OGC, HGC and CC ..................................................... 128

Figure 5.20 Comparison of the tested strength reduction factor with existing models ....... 129

Figure 5.21 Comparison of the measured thermal conductivity with existing model ......... 130

Figure 6.1 Elevation and cross section of the columns ....................................................... 135

Figure 6.2 Preparation of RC columns ............................................................................... 136

Figure 6.3 Stress-strain curve for reinforcing steel ............................................................. 137

Figure 6.4 Fabricated RC columns, (a) after casting, (b) at testing day .............................. 138

Figure 6.5 Test set-up ......................................................................................................... 140

Figure 6.6 Failure mode of OGC column ........................................................................... 141

Figure 6.7 Failure mode of HGC column ........................................................................... 142

Figure 6.8 Axial displacement versus axial load ................................................................ 143

Figure 6.9 Mid-height deflection versus axial load ............................................................ 144

Figure 6.10 Adopted material behaviour for cementitious grout ........................................ 146

Figure 6.11 Comparison between measured and predicted Load-Deflection curve for OGC

and HGC columns ............................................................................................................... 148

Figure 6.12 Comparison between measured and predicted Load-Axial displacement curve

for OGC and HGC columns ................................................................................................ 148

Figure 6.13 Predicted Load-Deflection and Load-Axial displacement curves for CC columns

............................................................................................................................................ 149

Figure 6.14 Typical FE analysis ......................................................................................... 150

Figure 6.15 Axial load versus mid-height deflection curves for OGCI-1 ........................... 152

Figure 6.16 Axial load versus mid-height deflection curves for OGCI-4 ........................... 153

Figure 7.1 Detailed dimensions of the reinforced column specimens................................. 158

Figure 7.2 Fire test set-up ................................................................................................... 160

Figure 7.3 OGC-1a column during the fire test at different times ...................................... 163

Figure 7.4 OGC-1b column during the fire test at different times. ..................................... 163

Figure 7.5 OGC-2 column during the fire test at different times ........................................ 164

Figure 7.6 HGC-1a column during the fire test at different times ...................................... 164

Figure 7.7 HGC-1b column during the fire test at different times ...................................... 165

Figure 7.8 HGC-2 column during the fire test at different times ........................................ 165

Figure 7.9 CC-1 column during the fire test at different times ........................................... 166

Page 15: Mechanical, Thermal and Structural Investigations of One

xiv

Figure 7.10 CC-2a column during the fire test at different times. ...................................... 166

Figure 7.11 CC-2b column during the fire test at different times ....................................... 167

Figure 7.12 CC-3 column during the fire test at different times ......................................... 167

Figure 7.13 A general view of the OGC columns after test ................................................ 168

Figure 7.14 A general view of the HGC columns after test ................................................ 168

Figure 7.15 A general view of the CC columns after test ................................................... 169

Figure 7.16 Different views of OGC columns after failure ................................................ 169

Figure 7.17 Different views of HGC columns after failure ................................................ 170

Figure 7.18 Different views of CC columns after failure ................................................... 170

Figure 7.19 Temperature T versus time t curves of OGC columns ..................................... 172

Figure 7.20 Temperature T versus time t curves of HGC columns ..................................... 173

Figure 7.21 Temperature T versus time t curves of CC columns ........................................ 173

Figure 7.22 Effect of concrete type on the temperature distribution of column section at

different points. ................................................................................................................... 174

Figure 7.23 Axial deformation versus time curves for OGC, HGC and CC columns ......... 176

Figure 7.24 Comparison of axial deformation versus time curves for the columns with

similar initial load levels. .................................................................................................... 176

Figure 7.25 Thermal conductivity of steel as a function of temperature ............................. 179

Figure 7.26 Specific heat of steel as a function of temperature .......................................... 180

Figure 7.27 Concrete tensile reduction factor versus temperature ...................................... 182

Figure 7.28 Finite element model for stress analysis .......................................................... 185

Figure 7.29 Tested versus measured Temperature as a function of Time for OGC columns

............................................................................................................................................ 187

Figure 7.30 Tested versus measured Temperature as a function of Time for HGC columns

............................................................................................................................................ 187

Figure 7.31 Tested versus measured Temperature as a function of Time for CC columns . 188

Figure 7.32 Comparison of axial deformation for OGC columns ....................................... 189

Figure 7.33 Comparison of axial deformation for HGC columns ....................................... 190

Figure 7.34 Comparison of axial deformation for CC columns .......................................... 190

Page 16: Mechanical, Thermal and Structural Investigations of One

Chapter 1 Introduction

1

Chapter 1 Introduction

General

This chapter will provide an overview of this thesis on the development of novel

one-part geopolymer mixes and investigation of their mechanical, thermal and

structural performance. It also includes the overview and significance, research

objective and plans, and the layout of this thesis.

Overview and Significance

Concrete is one of the major construction materials vastly utilised worldwide. This

extensive use of concrete has been associated with 5 to 7% of the global man-made

carbon dioxide emission over the last decades (Lloyd and Rangan 2010, Part, Ramli

et al. 2015). The environmental issues attributed to the production of ordinary

Portland cement (OPC) concrete inspired various research works aiming to decrease

its global carbon footprint. These attempts are varying from using supplementary

cementitious materials (SCMs) as partial cement replacement (Kroehong, Sinsiri et

al. 2011, Nath and Sarker 2011, Cheah and Ramli 2012, Cheah and Ramli 2013) to

developing a whole new cement-less binder, called geopolymers (Davidovits 1989,

Hardjito and Rangan 2005, Provis and Van Deventer 2009). The term “geopolymer”

was introduced in the 1970s by the French scientist Davidovits (Davidovits 1989).

Geopolymers are a member of inorganic polymers family and synthetised by

combining any pozzolanic compound or source of silica and alumina in the presence

of strong alkali activators (Davidovits 1994). The polymerisation process is divided

into three main steps: (1) Dissolution of Si and Al atoms from the source material

Page 17: Mechanical, Thermal and Structural Investigations of One

Chapter 1 Introduction

2

through the action of hydroxide ions under highly alkaline condition; (2)

transportation or orientation or condensation of precursor ions into monomers; (3)

Setting or polycondensation / polymerisation of monomers into 3D network of silico-

aluminates structures. The sialate (silicon-oxo-aluminate) network consists of SiO4

and AlO4 tetrahedra linked alternatively by sharing all oxygen atoms. Positive ions

(Na+, K+, Li+, Ca++, Ba++, NH4+, H3O

+) must be present in the formwork cavities to

balance the negative charge of Al3+. Poly(sialates) have the following empirical

formula proposed by Davidovits (Davidovits 1994):

𝑀𝑛{−(𝑆𝑖𝑂2)𝑧 − 𝐴𝑙𝑂2}𝑛,𝑤𝐻2𝑂 (1-1)

where M is cation such as potassium, sodium or calcium, n is the degree of

polycondensation and z is 1, 2, 3 (Davidovits 1994). The amorphous to semi-

crystalline three dimensional aluminosilicate structures can take one of the three

basic forms as shown in Figure 1.1:

Poly(sialate)

(-Si-O-Al-O-)

Poly(sialate-siloxo)

(-Si-O-Al-O-Si-O-)

Poly(sialate-disiloxo)

(-Si-O-Al-O-Si-O-Si-O-)

Figure 1.1 Types of polysialate (Davidovits 1994)

The availability of abundant industrial by-products such as fly ash, ground

granulated blast furnace slags (GGBFS), palm oil fuel ash (POFA), rice husk ash

(RHA), red mud (RM) etc which have been proved to have the potential to be used as

Page 18: Mechanical, Thermal and Structural Investigations of One

Chapter 1 Introduction

3

geopolymer source materials, supported the concept of OPC binders replacement

with geopolymers (Swanepoel and Strydom 2002, Bakharev 2005, Kumar, Kumar et

al. 2010, He, Jie et al. 2013, Salih, Ali et al. 2014). This reuse opportunity of

industrial by-products, also help to reduce required storage for disposal purposes

(McLellan, Williams et al. 2011). Furthermore, it has been found that geopolymers

indicate high compressive strength, negligible drying shrinkage, low creep, good

bond with reinforcement, and good resistance to acid, sulfate and fire and the

performance of geopolymer concrete structural members such as beams and columns

similar to that of OPC concrete members (Nematollahi, Sanjayan et al. 2015, Pan,

Tao et al. 2017).

However, the application of geopolymers so far has been limited to small scale and

in order to take the advantage of their superior environmental friendliness, the large

scale application of geopolymer should be promoted by overcoming several

drawbacks associated with the production of geopolymers. The conventional

geopolymer is made of a two-part mix, including alkaline solutions and solid

aluminosilicate precursor (Duxson, Fernández-Jiménez et al. 2007, Provis and Van

Deventer 2009). Handling large amounts of viscous, corrosive and hazardous alkali

activator solutions prohibits the mass production of geopolymers. Recently, the

development of one-part “just add water” geopolymers, in which the solid activators

are used has enhanced their commercial feasibility while their production is identical

to that of OPC mixes (Hajimohammadi, Provis et al. 2008, Nematollahi, Sanjayan et

al. 2015, Hajimohammadi and van Deventer 2017). Furthermore, it is well known

that heat treatment is essential for conventional geopolymer to achieve comparable

strength to that of OPC concrete (Ryu, Lee et al. 2013, Ranjbar, Mehrali et al. 2014)

which limits the widespread application of geopolymers. In order to eliminate the

Page 19: Mechanical, Thermal and Structural Investigations of One

Chapter 1 Introduction

4

heat curing requirement, the incorporation of various additives to the mixes such as

slag, high calcium fly ash, calcium hydroxide and OPC have been found to be useful

(Palomo, Fernández-Jiménez et al. 2007, Chindaprasirt, Chareerat et al. 2010,

Canfield, Eichler et al. 2014, Nath and Sarker 2015, Aliabdo, Elmoaty et al. 2016,

Assi, Ghahari et al. 2016).

Fire represents one of the most severe environmental conditions that structure might

experience and appropriate fire safety measurements are essential in the design of

high-rise buildings and infrastructure where concrete is mostly used as main

structural material due to its excellent fire resistance. However, fire causes

deterioration of concrete properties which is a significant reduction in strength,

stiffness and durability due to the physicochemical changes in the materials during

heating. Rapid heating due to fire can lead to large volume changes due to thermal

dilatations, and also thermal shrinkage and creep related to water loss. Large internal

stresses result in cracking of concrete or even large fractures. Furthermore, spalling

of concrete may occur in the case of rapid fire heating. Accordingly, the weak parts

of structural members should be identified and should be designed for the highest

possible level of safety. The heat transfer rate of materials is associated with their

thermal conductivity and heat capacity. Clearly, the generic information available on

the properties of OPC concrete subjected to fire is seldom applicable in case of

design of geopolymer concrete members. Therefore, the investigation of geopolymer

materials behaviour as a new development and their structural performance while

exposed to fire should be investigated comprehensively.

This research is a part of the ongoing research works in the area of development and

investigation of geopolymer concrete for structural applications subjected to fire. For

Page 20: Mechanical, Thermal and Structural Investigations of One

Chapter 1 Introduction

5

this purpose, material, thermal and microstructural properties of numerous developed

one-part geopolymers were evaluated. In the next step, the best mixes were selected

for production of reinforced concrete columns and their structural applications in

ambient and elevated temperatures were investigated. The results of the experimental

tests were compared with OPC concrete columns to assess the efficiency and

suitability of these novel materials as a practical solution.

Research Objectives and Scope of Work

The main objective of this research is to develop a commercially viable geopolymer

concrete mix and to examine mechanical, thermal and structural properties in

comparison with that of OPC concrete. Furthermore, the fire resistance of reinforced

geopolymer concrete columns were evaluated experimentally and compared with

those of reinforced OPC concrete columns. To achieve the above aims, the specific

objectives and scope of work are described as follows:

(1) Geopolymer concrete made with different activators and source materials at

ambient temperature.

Objectives:

(a) Evaluate the material properties of geopolymer concrete made with

different activators and their combinations and various source materials;

(b) Assess the stress-strain models of geopolymer concrete mixes.

Scope:

(a) Conduct experimental evaluation on the fresh properties of various

geopolymer mixes with different mix proportions;

(b) Conduct experimental investigation on the mechanical properties of

various geopolymer mixes with different mix proportions;

Page 21: Mechanical, Thermal and Structural Investigations of One

Chapter 1 Introduction

6

(c) Conduct microstructural analysis to examine the effects of different mix

proportions on the microstructure of geopolymer mixes.

(2) Application of reinforced geopolymer concrete columns.

Objectives:

(a) Investigate and evaluate the feasibility of using geopolymer concrete for

reinforced columns;

(b) Verify the feasibility of finite element (FE) prediction.

Scope:

(a) Conduct experimental investigation on reinforced concrete columns with

three different concrete mixes of hybrid OPC-geopolymer, one-part

geopolymer and OPC concrete mixes;

(b) Check whether finite element (FE) analysis modelling can be used for

prediction.

(3) Geopolymer concrete at elevated temperatures.

Objectives:

(a) Investigate and evaluate the compressive strength of geopolymer concrete

mixes at elevated temperatures;

(b) Investigate and evaluate the thermal properties of geopolymer concrete

mixes;

(c) Check whether current thermal models can be used for predication.

Scope:

(a) Conduct fire tests on concrete cylinders to investigate the influence of

different compositions on strength reduction at elevated temperatures and

then evaluate the test results by referring to current design guidelines and past

research findings;

(b) Conduct experimental investigation on the thermal conductivity and

specific heat of geopolymer mixes and then assess the results with design

codes and findings from other researchers.

Page 22: Mechanical, Thermal and Structural Investigations of One

Chapter 1 Introduction

7

(4) Fire resistance of reinforced columns made with hybrid OPC-geopolymer, one-

part geopolymer and OPC concrete.

Objective:

(a) Investigate and evaluate the fire resistance of reinforced columns made

with hybrid OPC-geopolymer, one-part geopolymer and OPC concrete.

Scope:

(a) Conduct experimental investigation on the fire resistance of hybrid OPC-

geopolymer and one-part geopolymer columns and compare the results with

those of OPC concrete;

(b) Check whether finite element (FE) analysis modelling can be used for

prediction.

Thesis Layout

This thesis is divided into eight main chapters as follows:

Chapter 1 has introduced the general background of geopolymers and fire hazards in

structures, then discussed the research motivations and proposed the study objectives

and scope.

Chapter 2 presents a comprehensive review on properties of geopolymer paste,

mortar and concrete at ambient and elevated temperatures, geopolymer concrete

members at ambient and elevated temperatures and current applications of

geopolymer concrete.

Chapter 3 presents experimental investigation on the fresh and hardened properties

of hybrid one-part OPC-geopolymer concrete together with reference mixes. The

slump, setting time, compressive strength and microstructure properties will be

assessed.

Page 23: Mechanical, Thermal and Structural Investigations of One

Chapter 1 Introduction

8

Chapter 4 presents experimental investigation on the fresh and hardened properties

of one-part geopolymer mixes. The slump, density, compressive strength and

microstructure properties will be assessed.

Chapter 5 presents experimental investigation on the mechanical and thermal

properties of the best hybrid OPC-geopolymer and one-part geopolymer concrete

mixes from Chapters 3 and 4. Mechanical properties including compressive

strength, modulus of elasticity (MOE), tensile strength (Brazil split), modulus of

rapture (MOR) and stress-strain curve and thermal properties including hot

compressive strength, Strain at peak stress, thermal conductivity and specific heat

will be measured and compared with those of OPC concrete results. Furthermore, the

compressive strength of all three types of concrete mixes at elevated temperatures is

evaluated. Also, the measured values of MOE, stress-strain curve, hot compressive

strength and thermal conductivity of mixes will be compared with those of existing

models.

Chapter 6 presents experimental investigation on reinforced concrete columns under

axial compression. Three different concrete mixes of hybrid OPC-geopolymer, one-

part geopolymer and OPC concrete are designed. The influence of mix composition

on the failure modes, development of axial and lateral strains, strength and ductility

of reinforced columns is discussed. The test results are compared with finite element

analysis results.

Chapter 7 presents experimental investigation on reinforced concrete columns

subjected to fire. Three different concrete mixes of hybrid OPC-geopolymer, one-

part geopolymer and OPC concrete are designed. This seeks the potential

Page 24: Mechanical, Thermal and Structural Investigations of One

Chapter 1 Introduction

9

improvement in fire resistance of geopolymer concrete columns. The test results are

compared with the developed finite element analysis results.

Chapter 8 draws the conclusions for the current research and outlines proposed

research works in geopolymer concrete area.

Page 25: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

10

Chapter 2 Literature Review

General

Chapter 1 has already introduced the background and motivation for this research

briefly. This chapter presents an overview of some basic concepts of geopolymers

including the geopolymers compounds and hydration products, introducing the

factors influencing the mechanical, microstructural and thermal properties of

geopolymers and recent developments in this area. It is followed by reporting the

previous findings regarding the structural performance of geopolymer concrete

members at ambient and elevated temperatures.

Source Materials

Geopolymer precursor could potentially be of any substance which is a source of

Silicon (Si) and Aluminium (Al). The nature of the aluminosilicate source

significantly affects the microstructure, mechanical, chemical and thermal properties

of geopolymers (Duxson, Fernández-Jiménez et al. 2007). Different studies have

been conducted to date to make geopolymers with various aluminosilicate precursor

type or combination of them.

Low calcium (ASTM Class F) fly ash has been considered as the most common

source material in the production of geopolymer, which is an industrial waste of

coal-burning power stations (van Jaarsveld and Van Deventer 1999, Swanepoel and

Strydom 2002, Hardjito and Rangan 2005, Wallah and Rangan 2006, Rattanasak and

Chindaprasirt 2009, Temuujin, Williams et al. 2009, Temuujin, van Riessen et al.

Page 26: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

11

2010, Ryu, Lee et al. 2013, Castel and Foster 2015). The general features of Class F

fly ash are illustrated in micrograph of Figure 2.1.

Figure 2.1 SEM micrograph of Class F fly ash particles (Jalal, Pouladkhan et al. 2015)

Ground granulated blast-furnace slag (GGBFS) which is an industrial waste of iron

production, could potentially be utilized to produce high strength geopolymers (Li

and Liu 2007, Kumar, Kumar et al. 2010, Oh, Monteiro et al. 2010, Parthiban,

Saravanarajamohan et al. 2013, Puligilla and Mondal 2013, Nath and Sarker 2014).

Figure 2.2 depicts the SEM micrograph of GGBFS.

Figure 2.2 SEM micrograph of GGBFS (Dave and Sahu 2013)

Page 27: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

12

The other aluminosilicate sources which have been used in the production of

geopolymer are: metakaolin which is a common industrial mineral (Duxson, Lukey

et al. 2007, Kong, Sanjayan et al. 2007, Bell, Driemeyer et al. 2009) (see Figure 2.3),

palm oil fuel ash (POFA) from the oil palm industry to generate electricity (Mijarsh,

Johari et al. 2014, Yusuf, Johari et al. 2014) , rice husk ash (RHA) from the rice

milling industry (Kusbiantoro, Nuruddin et al. 2012), red mud (RM) from the

alumina refining industry (Hajjaji, Andrejkovičová et al. 2013, He, Jie et al.

2013),etc.

Figure 2.3 SEM micrograph of metakaolin (Mansour, Abadlia et al. 2010)

The properties and hydration products of geopolymers are significantly affected by

the type and combination of source materials. The geopolymeric product of systems

with fly ash and metakaolin as source materials are mainly sodium aluminosilicate

hydrate gel while in slag based geopolymers mostly calcium silicate hydrate gel is

achievable (Singh, Ishwarya et al. 2015). The metakaolin-based geopolymer

production is consistent and properties during preparation and also development

could be controlled. However, their plate-shaped particles, result in increasing both

the difficulty of processing and the water demand of the system and accordingly

Page 28: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

13

rheological issues (Li, Sun et al. 2010). The better performance in terms of durability

and strength is obtained by replacing metakaolin with fly ash. The inclusion of slag

in geopolymers causes higher strength and acid resistance compared to those of

metakaolin and fly ash-based systems (Singh, Ishwarya et al. 2015).

Alkali Activators

The geopolymerisation process is significantly influenced by the type, dosage and

combination of alkali activators. In order to achieve desirable mechanical properties

for geopolymer, generally a strong alkaline medium is required (De Vargas, Dal

Molin et al. 2011) which causes transformation of glassy structure to a partially or

totally compacted composite (Khale and Chaudhary 2007).

Sodium and potassium based alkalis such as sodium hydroxide (NaOH), sodium

silicate (Na2SiO3), sodium carbonate (Na2CO3), sodium sulphate (Na2SO4),

potassium carbonate (K2CO3), potassium hydroxide (KOH), potassium sulphate

(K2SO4) have been mainly utilised to produce geopolymers (van Jaarsveld and Van

Deventer 1999, Xu and Van Deventer 2000). Conventionally, geopolymers are

produced from two-part mixing, i.e., alkaline solution activators are added to solid

aluminosilicate.

The combination of activators such as sodium hydroxide solution with sodium

silicate solution with different Na2SiO3/NaOH molar ratios has been studied in

different studies (Hardjito, Wallah et al. 2004, Olivia and Nikraz 2012, Sarker,

Haque et al. 2013). It was reported that by mixing the activators such as sodium

silicate and sodium hydroxide, better geopolymeric reaction among precursors and

solutions was observed (Rangan, Hardjito et al. 2005). Also, the more NaOH gel in

the presence of solid reactive materials, the more silicate and aluminate monomers

Page 29: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

14

are released (Khale and Chaudhary 2007). Accordingly, the proper concentration of

NaOH and Na2SiO3/NaOH molar ratios, and also alkaline activator/binder ratios, as

important factors influencing geopolymerisation development, have been

comprehensively investigated (Somna, Jaturapitakkul et al. 2011, Sukmak,

Horpibulsuk et al. 2013, Görhan and Kürklü 2014).

However, handling large volume of highly corrosive and viscous alkaline solutions

prevents the widespread field application of geopolymers. Also, controlling the ratio

of alkali to available silica and water and alkali content as important factors affecting

geopolymerisation is challenging in practice (Nematollahi, Sanjayan et al. 2015).

Therefore, there has been a tendency from researchers towards using solid alkaline

activators like sodium silicate, sodium hydroxide, their combination, etc

(Hajimohammadi, Provis et al. 2008, Yang, Song et al. 2008, Yang and Song 2009,

Nematollahi, Sanjayan et al. 2015). The feasibility of producing geopolymers with

solid activators has offered a new class of geopolymers called: “one-part” which

could be fabricated in an identical manner to that of OPC material. This development

encourages the large scale application of geopolymers in building industry.

Curing Conditions

Curing condition is another critical parameter that controls the geopolymer

properties. It is well known that heat curing is one of the requirements for low

calcium geopolymers to set and achieve reasonable early strength (Rangan 2008,

Lloyd and Rangan 2010, Ryu, Lee et al. 2013). Heat curing accelerates the reaction

process by accelerating the silica and alumina release. The geopolymerisation

process could be extremely improved by exposing the mix to temperature range of 60

℃ to 90 ℃ for 8-24 hour (Singh, Ishwarya et al. 2015). On the other hand, curing at

Page 30: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

15

very high temperature for longer duration of time weakens the geopolymer structure

and cause cracking and decrease in strength (Khalil and Merz 1994, Van Jaarsveld,

Van Deventer et al. 2002). Therefore, the curing regime and period for geopolymeric

materials should be appropriately taken into consideration to obtain desirable

mechanical properties (Van Jaarsveld, Van Deventer et al. 2002).

Numerous research works have been conducted with the aim of removing necessity

of heat curing for geopolymers. In most of these studies, positive effect of the

addition of various additives such as slag, metakaolin, OPC, etc. as source for

calcium have been reported (He, Zhang et al. 2012, Nath and Sarker 2012, Deb, Nath

et al. 2014, Nath and Sarker 2014, Nath and Sarker 2015). Despite the fact that

inclusion of calcium compounds might interfere with geopolymerisation, it enhances

the process by reducing the setting time and increasing the early strength (Temuujin,

Van Riessen et al. 2009).

Behaviour of Geopolymer Concrete at Ambient

Temperature

Fresh properties

The workability of geopolymers is substantially influenced by: concentration of

activators, workability aids, water to geopolymer solid ratios, type of aluminosilicate

sources, etc (Pacheco-Torgal, Labrincha et al. 2014). The fly ash based geopolymers

developed by Hardjito et al. (Hardjito, Wallah et al. 2004) showed slump variation in

the range of 100 to 250 mm for alkaline molarity range of 8 to 14. Although by

increasing the water to binder ratio better workability is achievable, the compressive

strength is decreased significantly.

Page 31: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

16

SiO2/Al2O3 ratio has a considerable influence on the setting time of geopolymers and

its increase causes an increase of setting time (De Silva, Sagoe-Crenstil et al. 2007,

De Vargas, Dal Molin et al. 2011). Furthermore, increasing the temperature causes

faster setting of low calcium geopolymers due to quicker release of Si and Al from

source materials (Pacheco-Torgal, Labrincha et al. 2014). The addition of calcium

compounds in low calcium geopolymer systems, significantly decreases the setting

time (Xu and Van Deventer 2000, Yip, Lukey et al. 2005).

Mechanical properties

The mechanical properties of geopolymer concrete are influenced by many factors

such as type and combination of source materials and activators, mix proportions and

curing regime (Pacheco-Torgal, Labrincha et al. 2014). For instance, SiO2/M2O (M=

Na+ or K+) ratio has a substantial effect on dissolution extent of species in alkali

activators while by increasing the silica content of mixes, the rate of

geopolymerisation is decreased and early paste solidification before completion of

reaction is observed (Duxson, Provis et al. 2005, Duxson, Fernández-Jiménez et al.

2007). The relationship between SiO2/M2O ratio and compressive strength of

geopolymers show that either by increasing M2O or decreasing SiO2, improvement

of compressive strength is achieved (van Jaarsveld and Van Deventer 1999, Xu and

Van Deventer 2002). Khale and Chaudhary (Khale and Chaudhary 2007) suggested

the following composition ranges to develop a reasonable geopolymer structure:

M2O/SiO2: 0.2-0.48, SiO2/Al2O3: 3.3-4.5, H2O/M2O: 10-25 and M2O/Al2O3: 0.8-1.6.

Based on various studies on geopolymer concrete, the compressive strength (𝑓𝑐′) is

ranging between 30 to 80 MPa (Hardjito, Wallah et al. 2004). An ultra-high-

performance geopolymer concrete with strength up to 175 MPa was developed in

which the binder included up to 33% fly ash and the remaining was replaced by slag

Page 32: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

17

and silica fume (Ambily, Ravisankar et al. 2014). Also, a high strength of 120 MPa

was achieved for geopolymer mortars in which NaOH with molarity of 14% was

used and mix was cured at a temperature of 115 ℃ for 24 hrs (Atiş, Görür et al.

2015).

The modulus of elasticity (MOE) of heat cured geopolymer concrete is lower than

that of OPC concrete. However, the ambient cured geopolymers (with inclusion of

calcium compounds) showed identical MOE in comparison with OPC concrete (Nath

and Sarker 2017).

Geopolymer concrete has relatively low splitting tensile (𝑓𝑐𝑡) and flexural strength

(𝑓𝑐𝑡,𝑓′ ) which is similar to OPC concrete. With similar compressive strength, heat

cured geopolymer has higher 𝑓𝑐𝑡 than that of OPC while ambient cured geopolymer

has similar value as OPC concrete (Deb, Nath et al. 2014). However, both heat and

ambient cured geopolymers have higher 𝑓𝑐𝑡,𝑓′ than that the OPC concrete with similar

𝑓𝑐′ (Deb, Nath et al. 2014, Nath and Sarker 2017).

Microstructure of geopolymers

Many factors are affecting the microstructure of geopolymers such as: the nature of

aluminosilicates, chemical composition and processing condition (Duxson,

Fernández-Jiménez et al. 2007, Khale and Chaudhary 2007). Different techniques

such as scanning electron microscope (SEM), energy dispersive spectroscopy (EDS),

Infrared Spectra (IR), x-ray diffraction (XRD), nuclear magnetic resonance (NMR),

etc have been used to investigate the microstructure properties of geopolymers.

IR and NMR have been found to be suitable tools to characterise geopolymer

materials (Van Jaarsveld, Van Deventer et al. 1999). XRD pattern and IR spectrum

Page 33: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

18

of a typical fly ash based geopolymer are shown in Figures 2.4 and 2.5, respectively.

As seen in Figure 2.4, the substantial part of structure contains amorphous phases

between 20º and 40º 2𝜃 with minor crystalline phases related to the source materials

(Komnitsas and Zaharaki 2007).

Figure 2.4 XRD pattern of fly ash based geopolymer (M=mullite, Q=quartz) (Škvára,

Kopecký et al. 2006)

As seen in Figure 2.5, fly ash and fly ash based geopolymer show different IR

spectrum patterns. The major aspect of IR spectra within the range of 1080 and 1090

cm-1 in pure fly ash shifts toward lower wavelength after reaction. This speak is

associated with the Si-O-Si or Al-O-Si symmetric stretching mode and the shift

toward lower values shows the extent of reaction (Van Jaarsveld, Van Deventer et al.

1999, Škvára, Kopecký et al. 2006).

Page 34: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

19

Figure 2.5 IR spectrum of fly ash and fly ash based geopolymer (Škvára, Kopecký et al.

2006)

By adding calcium compound such as slag to fly ash based geopolymer, major

changes in microstructure is observed. Figure 2.6 depicts the XRD pattern of

geopolymers with varying slag content. All the spectra regardless of slag

concentration have crystalline peaks of mullite and quartz from the original raw fly

ash. By increasing the slag content, the intensity of crystalline peaks decreases. No

new crystalline peak upon activation of slag was observed which shows that

crystalline CSH was not a dominant product. Amorphous CSH within a

geopolymeric gel (calcium based geopolymer) was suggested to be part of the

hydration product.

Page 35: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

20

Figure 2.6 XRD pattern of geopolymers with varying slag content (Kumar, Kumar et al.

2010)

The use of SEM and EDS has also been useful to provide qualitative and quantitative

information regarding the changes in microstructure of geopolymers with different

mix designs (Bakharev 2005, Duxson, Provis et al. 2005, Yip, Lukey et al. 2005,

Škvára, Kopecký et al. 2006, Temuujin, Williams et al. 2009, Somna, Jaturapitakkul

et al. 2011, Ryu, Lee et al. 2013, Nath and Sarker 2014, Suwan and Fan 2014, Nath,

Maitra et al. 2016, Hajimohammadi and van Deventer 2017).

Properties of Geopolymers at Elevated Temperatures

Spalling

Spalling is the breaking away of concrete layers due to the build-up pore pressure or

thermal stress (Ali, Sanjayan et al. 2017). Generally, geopolymer concrete exhibits

Page 36: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

21

less spalling in comparison with OPC concerte. Zhao and Sanjayan (Zhao and

Sanjayan 2011) compared the behaviour of geopolymer and OPC concrete cylinders

(𝑓𝑐′=40-100 MPa) subjected to elevated temperatures with two different heating rate

regimes. Regardless of the heating rate and strength, no spalling was observed for

geopolymer specimens while OPC concrete samples experienced minor to severe

spalling. This excellent behaviour of geopolymer concrete is due to its porous

structure which provides the rapid release of steam from the matrix and hence causes

lower internal pore pressure.

Sarker et al. (Sarker, Kelly et al. 2014) also tested the behaviour of geopolymer and

OPC concrete specimens exposed to ISO 834 standard fire. They reported severe

spalling for OPC specimens at 800 and 1000℃, whereas no spalling was occurred for

geopolymer concrete samples.

In another research attempt by Kong and Sanjayan (Kong and Sanjayan 2010),

geopolymer concrete mixes containing aggregate size <10 mm experienced

considerable spalling and cracking while the specimens with aggregate size >10 mm

showed more steady behaviour. This behaviour was also observed by Pan and

Sanjayan (Pan and Sanjayan 2010) while explosive spalling at high temperatures was

observed when maximum aggregate size was below 10 mm regardless of mix

composition.

Mechanical properties

Based on Eurocode 2, OPC concrete experiences strength loss after exposing to

elevated temperatures. However, for heat cured geopolymer concrete, an increase of

compressive strength has been observed due to further geopolymerisation (Shaikh

Page 37: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

22

and Vimonsatit 2015). For ambient cured geopolymer, the strength retention might

be better in comparison with OPC concrete samples (Cao, Tao et al. 2018).

Kong and Sanjayan (Kong and Sanjayan 2008) measured the compressive strength

loss of geopolymer paste and concrete exposed to temperatures up to 800 ℃.

Geopolymer paste and concrete showed different behaviour when exposed to

elevated temperatures with similar heating regime. The strength of geopolymer paste

increased after thermal exposure while on the other hand, strength decreased for

geopolymer concrete. The strength loss of geopolymer concrete is due to thermal

incompatibility between geopolymer paste and aggregate while at elevated

temperatures aggregate exhibits expansion and geopolymer experiences contraction.

Mechanical properties of geopolymers in the hot state and also after cooling were

measured by Pan and Sanjayan (Pan and Sanjayan 2010). Within the range of 200-

290 ℃, further geopolymerisation caused an increase in strength, followed by further

increase due to thermal expansion up to 520 ℃ and reached 179% of the initial

strength and after this temperature remained almost constant. The glass transition

behaviour was observed at temperature of 560 ℃ for geopolymer samples. The

residual strength had the same pattern as hot strength but with loss of strength after

cooling due to thermal shock. The stress-strain behaviour of geopolymers exhibited

that at any temperature, a brittle type of failure was observed.

The geopolymer softening at high temperatures is affected by type of alkali activator

(Pan and Sanjayan 2012). The geopolymer softening was observed at 610 ℃

(±20 ℃) for sodium based geopolymers, 570 ℃ for sodium/potassium based mixes

and 800 ℃ for potassium based geopolymers.

Page 38: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

23

Pan et al. (Pan, Sanjayan et al. 2014) also measured the strength of geopolymer and

OPC paste and concrete exposed to temperatures up to 550 ℃. Geopolymer paste

showed different behaviour compared to geopolymer concrete. The strength for

geopolymer paste reached up to 192% of its initial strength at 550 ℃, while strength

lost was seen for geopolymer concrete. A higher strength loss was observed for

geopolymer concrete compared with OPC concrete during heating which is due to

experiencing large thermal incompatibility between aggregate and matrix in the

range of 200-450 ℃ as well as internal cracks and excessive damage due to absence

of transitional thermal creep in the range of 250-550 ℃.

The hot strength for geopolymer concrete mixes developed by Shaikh and

Vimonsatit (Ahmed and Vimonsatit 2012) decreased up to 400 ℃ due to thermal

differentials between coarse aggregate and paste. At 600 and 800 ℃, the hot strength

for geopolymer concrete was higher than that of OPC concrete. The prediction of

compressive strength up to 400 ℃, based on Eurocode EN 1994:2005, gave a good

agreement with experimental data, but after this range the strength was

underestimated.

Thermal properties

Bakharev (Bakharev 2006) evaluated the thermal properties of fly ash based

geopolymer concrete upon firing at 800-1200 ℃ while sodium and potassium based

alkaline activators were used. They found that thermal stability of mixes containing

sodium as activator was low and microstructure experienced major changes. The

quick degradation of strength at 800 ℃ due to significant increase of the average

pore size was seen. On the other hand, geopolymers containing potassium silicate as

activator exhibited better thermal stability and remained amorphous up to 1200 ℃.

Page 39: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

24

Pan et al. (Pan, Tao et al. 2018) investigated the thermal properties of low calcium

and high calcium geopolymer binders with different mix designs and source

materials at elevated temperatures up to 600 ℃. It was reported that by increasing

temperature, thermal conductivity of low calcium geopolymers increases while high

calcium systems experience decrease in thermal conductivity. In terms of specific

heat, by increasing temperature for both types of binders, a general increasing trend

is observed due to further geopolymerisation.

Recent Developments in Geopolymers

Blended geopolymers

Blended geopolymer is the new category of geopolymeric materials in which the

geopolymer precursor is a combination of more than one waste material. The heat

requirement and sensitivity of geopolymer properties to curing regime is considered

as one of the main barriers preventing the wide and large scale application of

geopolymers in engineering practice.

In order to improve the geopolymerisation at room temperature, researchers have

incorporated various additives, such as slag, high calcium fly ash, calcium hydroxide

and OPC, as additional calcium sources in geopolymer mixes. The calcium sources

were found to play an important role to shorten the setting time and also improve the

mechanical properties (Palomo, Fernández-Jiménez et al. 2007, Chindaprasirt,

Chareerat et al. 2010, Nath and Sarker 2012, Suwan and Fan 2014, Nath and Sarker

2015, Aliabdo, Elmoaty et al. 2016, Assi, Ghahari et al. 2016, Shehab, Eisa et al.

2016).

Page 40: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

25

The OPC blended geopolymers have been employed in several studies recently. For

example, Nath and Sarker (Nath and Sarker 2015) replaced up to 12% of fly ash by

OPC as a calcium silicate source. In this experiment, alkaline activator solution of

combined sodium silicate and sodium hydroxide was used, whilst the curing regime

was ambient curing for all specimens. The authors concluded that the optimum

mixture was achieved when OPC was used to replace 5% of the total binder and the

ratio of liquid activator to the binder was 0.4. The mixture was cured at ambient

temperature and set reasonably well compared to OPC concrete.

Suwan and Fan (Suwan and Fan 2014) evaluated the effects of OPC replacement and

mixing procedure on properties of geopolymers. The replacement ratio of fly ash

with OPC varied from 5% to 70%, and its influence on the setting time, compressive

strength and microstructure of the geopolymer was studied. Meanwhile, the sequence

of adding fly ash, sodium silicate and sodium hydroxide was changed in the

specimen preparation. Results showed that by increasing the amount of OPC from

5% to 70%, the setting time of geopolymer was significantly reduced from 24 h to

0.5 h while the early strength at 3 days increased from 0 to 21.91 MPa. It was also

reported that the mixing sequence could directly affect the properties of geopolymer.

A total of three mixing methods were adopted. In the first mixing method, fly ash

was initially mixed with sodium hydroxide solution for 90s before the sodium

silicate was added. In the second method, sodium silicate and sodium hydroxide

solutions were premixed and left over night and then fly ash was added during

mixing. In the last method, fly ash, sodium hydroxide and sodium silicate were dry-

mixed before adding water. It was found that only the third mixing sequence ensured

the setting of the mix within the first 24 h. However, the mixes prepared using the

Page 41: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

26

first and second methods showed higher early strength as a result of enhanced

chemical reaction.

In another study conducted by Shehab et al. (Shehab, Eisa et al. 2016) ], mechanical

properties of fly ash based geopolymer concrete mixes with a partial replacement of

OPC were investigated by varying the OPC replacement ratio and activator solution

concentration. The authors found that the mix with 50% OPC replacement had the

highest compressive strength, splitting tensile strength, bond strength and flexural

strength at 28 days.

One-part geopolymers

One-part “just add water” geopolymer mixes are a newly emerged type of

geopolymers, in which the ready-mix precursor of aluminosilicate source and solid

activator are directly mixed with water similar to production of OPC based concrete

(Yang, Song et al. 2008). The aim of developing one-part geopolymers is to promote

the large-scale application of geopolymer binders in building industry by avoiding

the difficulties of handling large quantities of highly corrosive and viscous alkaline

solutions in conventional geopolymer mixes. In this regard, there have been attempts

by researchers to develop one-part geopolymer mixes using different methods.

In case of one-part geopolymers, the phase assemblage of final products are

significantly affected by the choice of the silica feedstock and the initial SiO2 / Al2O3

ratio, resulting in the formation of either geopolymer-zeolite composite or to

virtually fully amorphous geopolymers (Sturm, Greiser et al. 2015, Sturm, Gluth et

al. 2016, Greiser, Gluth et al. 2018). Thus, selecting the silica starting material with

suitable dissolution kinetics considering the overall SiO2 / Al2O3 ratio of one-part

mixture, results in a formation of phase-pure sodium aluminosilicate gel with

Page 42: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

27

promising engineering properties such as high mechanical properties (Greiser, Gluth

et al. 2018).

In an early attempt, Koloušek et al. (Koloušek, Brus et al. 2007) synthetised one-part

geopolymer by direct calcination of low-quality kaolin with powdered hydroxide at

550 ℃ for 4 h. However, the mixes developed very low 7-day compressive strength

of only 1 MPa. Feng et al. (Feng, Provis et al. 2012) also used calcination method to

activate albite with sodium hydroxide or sodium carbonate at elevated temperatures

(850-1150 ℃ ) and the produced one-part geopolymer mix achieved 28-day

compressive strength of over 40 MPa. Subsequently, Ke et al. (Ke, Bernal et al.

2015) prepared one-part geopolymer mixes by calcination of red mud with sodium

hydroxide at 800 ℃ and their product gained compressive strength of 10 MPa after

28 days. However, the need for high temperature calcination in these studies to

obtain one-part geopolymer limits the application of these products in construction

industry.

Without calcination, Hajimohammadi et al. (Hajimohammadi, Provis et al. 2008)

directly mixed geothermal silica with solid sodium aluminate to produce one-part

geopolymer. The water content was found to play a vital role during

geopolymerisation, and samples with less water exhibited a more crystalised

structure. This research group also examined the effects of alumina release rate

(Hajimohammadi, Provis et al. 2010) and silica availability (Hajimohammadi, Provis

et al. 2011) on the mechanism of geopolymerisation of the developed one-part mixes.

However, these studies mainly focused on the chemistry and microstructure rather

than the mechanical properties of the one-part geopolymer mixes.

Page 43: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

28

There are also some attempts to develop one-part geopolymer mixes using common

sources of aluminosilicate, such as fly ash and slag. Yang et al. (Yang, Song et al.

2008) and Yang and Song (Yang and Song 2009) developed one-part geopolymer by

mixing fly ash or slag with sodium silicate powder or a combination of solid sodium

silicate and sodium hydroxide. For the pure fly ash and pure slag systems, the

achieved 28-day compressive strengths were 9.45 and 50 MPa, respectively. It was

found that a higher strength was achieved for the pure slag system associated with

the use of the highly hygroscopic and corrosive sodium hydroxide.

In contrast, Nematollahi et al. (Nematollahi, Sanjayan et al. 2015) synthesised one-

part geopolymers using different combinations of fly ash, slag and hydrated lime

activated by different grades of solid sodium silicate. The compressive strengths of

the mixes cured at ambient temperature varied between 5.2 and 37.3 MPa at 28 days,

whereas the compressive strength of the reference mix was 42.5 MPa using

conventional geopolymer technology. However, in this research the microstructure

properties of one-part geopolymers were not studied which is an important step

towards better understanding of their mechanical properties and structural

performance.

In terms of durability, one-part geopolymers made of silica and sodium aluminate

have been found to have resistance to sulfuric acid attack. However, addition of slag

in these binders caused a significant decrease with respect to sulfuric acid resistance

due to formation of expansive calcium sulfate phases (Sturm, Gluth et al. 2018).

Furthermore, the feasibility of using one-part geopolymers to develop cement-less

composites with excellent tensile ductility and material sustainability has been

proved (Nematollahi, Sanjayan et al. 2017, Nematollahi, Sanjayan et al. 2017).

Page 44: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

29

The feasibility of developing one-part geopolymer mixes encourages their

commercial workability. To achieve this aim, research efforts should be focused on

evaluating the structural performance of one-part geopolymer concrete members.

Performance of Geopolymer Concrete Members at

Ambient Temperature

Since the development of geopolymers, most of the studies have been focused on

micro-scaling while recent researchers have conducted several series of tests to

examine the structural performance of geopolymer concrete columns, beams, slabs,

etc. As an innovative building material, the structural performance of geopolymer

concrete members is one of the most crucial requirements to introduce such a

concrete for engineering practice applications. Furthermore, the behaviour of

geopolymer concrete members should meet the requirements of the existing design

provisions for OPC concrete members to ensure their safe and convenient use for

design of these novel members. Knowledge and findings from experimental and

numerical simulations, empirical formulations and appropriate assumptions provide

expertise for practicing engineers for a more effective and safer design (Mo,

Alengaram et al. 2016). In the following sections, a review of published research

studies regarding reinforced geopolymer concrete members is provided.

Geopolymer concrete beams

Sumajouw et al (Sumajouw and Rangan 2006) tested twelve simply supported

reinforced geopolymer concrete beams with various tensile steel reinforcement ratio

and concrete compressive strength. The details of beam specimens are shown in

Table 2.1 and the beams’ configuration after demoulding are depicted in Figure 2.7.

It should be noted that the beam specimens of this study were kept at room

Page 45: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

30

temperature after casting for 3 days before subjecting to heat curing at 60 ℃ for 24

hours.

In general, the performance of reinforced geopolymer concrete beams was identical

to that of OPC concrete beams with respect to effects of tensile reinforcement ratio

and compressive strength on cracking pattern flexural capacity and ductility index.

All beams collapsed in a ductile pattern followed by crushing of concrete in the

compression zone. The compressive strength increase caused an increase in the

cracking moment. The flexural capacity was substantially influenced by tensile

reinforcement ratio. The ductility of beams increased significantly by decreasing the

tensile reinforcement ratio to less than 2%. The comparison of flexural capacity of

geopolymer concrete beams with design provisions of AS 3600-2005 showed a good

correlation between the measured and calculated bending capacity results with the

average test-to-calculated ratio of 1.11.

Figure 2.7 Geopolymer concrete beams after demoulding (Sumajouw and Rangan 2006)

Page 46: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

31

Table 2.1 Beam details (Sumajouw and Rangan 2006)

Series Beam Beam

Dimension

(mm)

Reinforcement Tensile

Reinforcement

ratio (%)

Concrete

compressive

strength

(MPa)

Compression Tension

Ι GBI-1 200X300X3300 2N12 3N12 0.64 37

GBI-2 200X300X3300 2N12 3N16 1.18 42

GBI-3 200X300X3300 2N12 3N20 1.84 42

GBI-4 200X300X3300 2N12 3N24 2.69 37

ΙΙ GBII-1 200X300X3300 2N12 3N12 0.64 46

GBII-2 200X300X3300 2N12 3N16 1.18 53

GBII-3 200X300X3300 2N12 3N20 1.84 53

GBII-4 200X300X3300 2N12 3N24 2.69 46

ΙΙΙ GBIII-1 200X300X3300 2N12 3N12 0.64 76

GBIII-2 200X300X3300 2N12 3N16 1.18 72

GBIII-3 200X300X3300 2N12 3N20 1.84 72

GBIII-4 200X300X3300 2N12 3N24 2.69 76

Yost et al. (Yost, Radlińska et al. 2013) explored the behaviour of reinforced alkali

activated fly ash concrete beams subjected to four-point loading. A total of eighteen

beam specimens, including nine geopolymer concrete beams and nine control OPC

concrete beams, were tested with three various beam designs of under-reinforced

(U), over-reinforced (O) and shear critical (S). The test samples are illustrated in

Figure 2.8.

Page 47: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

32

Figure 2.8 Test sample detail of beams (Yost, Radlińska et al. 2013)

In general, the performance of geopolymer and OPC concrete beams in terms of

cracking width and pattern, neutral axis location and flexural and shear capacity were

similar. However, for O type geopolymer concrete beams, load-deflection response

was perfectly linear to failure while slight nonlinear behaviour was observed for

companion OPC concrete beams. In flexure, the geopolymer concrete beams were

more brittle material at ultimate than OPC concrete beams. Overall, geopolymer

concrete flexural members can be analysed and designed according to ACI 318-08

procedure established for OPC concrete beams.

Dattatreya et al. (Dattatreya, Rajamane et al. 2011) fabricated a total of eighteen

beams made of three OPC concrete and six geopolymer concrete mixes with varying

compressive strength between 17 to 63 MPa. The beams were designed with 1.82 to

Page 48: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

33

3.33% tension reinforcement and having various ranges of fly ash and slag

combination in binder.

The authors concluded that load-deflection behaviour, the crack pattern and failure

modes of the OPC and geopolymer concrete beams were identical. However, the

cracking moment and service load moment of geopolymer concrete beams were

slightly lower than that of OPC concrete beams. The peak load value for OPC and

slag based geopolymer concrete beams were similar while geopolymer concrete

beams showed higher values at the same or lesser moment. The cracking, service and

ultimate moment carrying capacity of the geopolymer concrete beams, calculated

according to OPC concrete provisions and strain compatibility approach exhibited

acceptable correlation. Furthermore, it was reported that the computational

procedures of evaluating the performance parameters of the OPC beams can be used

in case of geopolymer concrete beams, but the prediction might not be conservative

in all the cases. The authors concluded that although there is a fair agreement

between the measured and predicted (based on IS 456:2000) deflection of reinforced

geopolymer concrete beams, still further studies are required to predict the structural

behaviour of geopolymer concrete beams.

In several studies, researchers conducted numerical analysis using ANSYS program

to evaluate the behaviour of geopolymer concrete beams and found reasonable

agreement between the predicted and experimental results (Kumaravel and

Thirugnanasambandam 2013, Kumaravel, Thirugnanasambandam et al. 2014). In a

recent study, Nguyen et al. (Nguyen, Le et al. 2016) utilized ABAQUS software to

validate the experimental results of heat cured reinforced geopolymer concrete

beams. The configuration of beam model in ABAQUS is shown in Figure 2.9. The

Page 49: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

34

finite element simulation with ABAQUS software, showed marginal difference in

terms of deflection values, but still good agreement with respect to simulated load-

deflection. Therefore, the use of ANSYS and ABAQUS software could be beneficial

to evaluate the performance of geopolymer concrete members.

Figure 2.9 Reinforcement of beam in ABAQUS (Nguyen, Le et al. 2016)

The structural performance of reinforced concrete beams with different mix designs

and source materials have also been investigated. Andalib et al. (Andalib, Hussin et

al. 2014) incorporated palm oil fuel ash (POFA) in fly ash based geopolymer

concrete with various range of POFA-to-fly ash ratio to pour the reinforced beams

and evaluate their flexural behaviour. They concluded that these beams had similar

cracking pattern and ultimate moment to that of OPC concrete beams.

In another study, Kathirvel and Kaliyaperumal (Kathirvel and Kaliyaperumal 2016)

investigated the effect of inclusion of recycled concrete aggregate (RCA) in

reinforced geopolymer concrete beams. The capacity of the beams increased with the

increase in RCA content while the optimum results were achieved at 75%

Page 50: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

35

replacement level with 22.53% increase in the loading capacity. Furthermore, the

deflection and ductility behaviour of the geopolymer beams were improved by

increasing the RCA content. Design provisions of ACI 318-11 were found to yield

conservative prediction of the ultimate moment of the geopolymer concrete beams

for all concrete mixes.

The effect of glass fibre inclusion on the flexural behaviour of reinforced

geopolymer concrete beams was also investigated (Srinivasan, Karthik et al. 2014).

Glass fibre content of the mixes varied in the range of 0.01%, 0.02%, 0.03% and

0.04% of the total volume of concrete. The load carrying capacity and deflections of

all fibre reinforced geopolymer concrete beams were more than that of geopolymer

beams. The initial cracking load and ultimate load of fibre reinforced beams were 10-

30% and 12-35% more than geopolymer beams, respectively.

Geopolymer concrete columns

Several studies have been conducted to examine the behaviour of reinforced

geopolymer concrete columns, as crucial elements that carry axial loading.

Sumajouw (Sumajouw and Rangan 2006) fabricated twelve slender reinforced

geopolymer concrete with two reinforcement ratios of 1.47% and 2.95% with

strength grades of 40 and 60 MPa. The axial loads were applied with eccentricities

ranging from 15 to 50 mm. The details and set-up of the columns are depicted in

Figures 2.10 and 2.11 and details of load eccentricity, concrete properties and

reinforcement are shown in Table 2.2.

Page 51: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

36

Figure 2.10 Column details and set-up (Sumajouw and Rangan 2006)

Figure 2.11 Column in the test machine (Sumajouw and Rangan 2006)

Page 52: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

37

Table 2.2 Column details (Sumajouw and Rangan 2006)

Series Column Load

eccentricity

(mm)

Concrete

compressive

strength (MPa)

Longitudinal

reinforcement

bars Ratio

I OGCI-1 15 42 4N12 1.47

OGCI-2 35 42 4N12 1.47

OGCI-3 50 42 4N12 1.47

II OGCII-1 15 43 8N12 2.95

OGCII-2 35 43 8N12 2.95

OGCII-3 50 43 8N12 2.95

III OGCIII-1 15 66 4N12 1.47

OGCIII-2 35 66 4N12 1.47

OGCIII-3 50 66 4N12 1.47

IV OGCIV-1 15 59 8N12 2.95

OGCIV-2 35 59 8N12 2.95

OGCIV-3 50 59 8N12 2.95

The crack pattern and failure modes of geopolymer concrete columns were identical

to those of OPC concrete columns. The crack originated at column mid-height, then

continued along the length of the column as shown in Figure 2.12. The failure mode

was flexural and longitudinal reinforcement buckled in the compression zone, mostly

at low eccentricity level. The column’s ultimate capacity was affected by the steel

reinforcement ratio, concrete compressive strength and load eccentricity. The load-

carrying capacity increased by decreasing the load eccentricity and increasing the

reinforcement ratio and concrete compressive strength.

Sumajouw (Sumajouw, Hardjito et al. 2007) compared the ultimate capacity of

geopolymer concrete columns with the design provisions of AS 3600-2004 and ACI

318-02. The measured and predicted load capacity of columns were in good

Page 53: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

38

agreement with a mean test-to-predicted ratio of 1.03, which shows these provisions

are applicable to reinforced geopolymer concrete columns.

This numerical study also indicated that the columns with smaller load eccentricity,

higher concrete strength and higher reinforcement ratio failed in a brittle manner.

However, the mid-height deflection of columns increased with the increase in the

load eccentricity, decrease in concrete strength and reinforcement ratio as shown in

Figure 2.13.

Figure 2.12 Failure mode of geopolymer concrete columns (Sumajouw and Rangan 2006)

Page 54: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

39

Figure 2.13 Load-deflection curves for different tested geopolymer concrete columns

(Sumajouw, Hardjito et al. 2007)

Later on, Sarker (Sarker 2009) conducted numerical analysis on the results of the

reinforced geopolymer concrete columns tested by Sumajouw (Sumajouw, Hardjito

et al. 2007) to develop an appropriate model to predict the load-deflection

performance and load-carrying capacity of geopolymer concrete members. The

modified stress-strain formulation suggested by Popovics (Popovics 1973) for OPC

concrete were used in a non-linear analysis of reinforced concrete columns. The

calculated values of the ultimate loads and mid-height deflection using the non-linear

analysis of columns agreed reasonably well with the test results with the mean value

of test-to-prediction ratio of 1.03 and 1.14, respectively.

Sujatha (Sujatha, Kannapiran et al. 2012) fabricated a total of twelve slender circular

columns including six heat cured geopolymer concrete columns and six OPC

concrete columns. The columns configuration is illustrated in Figure 2.14. Both the

geopolymer and OPC concrete designed to achieve two grades of 30 and 50 MPa.

The geopolymer concrete columns showed identical performance in comparison with

OPC concrete columns. The geopolymer concrete columns showed 34% higher load

Page 55: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

40

carrying capacity and also greater rigidity but slightly lesser deflection compared to

the companion OPC concrete columns.

Figure 2.14 Configuration of circular geopolymer concrete columns (Sujatha, Kannapiran et

al. 2012)

Ganesan et al. (Ganesan, Abraham et al. 2015) evaluated the inclusion of steel fibres

on performance of geopolymer columns with 2.01% reinforcement ratio and

subjected to monotonic axial compression. The test variables included varying ranges

of steel fibres (up to 1.0%) and aspect ratio (I/d) of slender columns.

The ultimate capacity of the geopolymer concrete columns was improved up to 56%

due to fibre-bridging effect which restricted early concrete cover spalling.

Furthermore, the inclusion of steel fibres caused enhancement in ductility (up to 29%

Page 56: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

41

increase) of geopolymer concrete columns confirmed by the increase of strain at the

peak axial compressive stress and area under the stress-strain curve.

The performance of geopolymer concrete columns subjected to combined axial load

and biaxial bending were evaluated by Rahman and Sarker (Rahman and Sarker

2011). The concrete strength varied in the range of 37 to 63 MPa and the

reinforcement ratio was 1.47% or 2.95%. The description of column samples is

shown in Table 2. 3 and the column test set-up is depicted in Figure 2.15.

Table 2.3 Test variables of the columns (Rahman and Sarker 2011)

Column Reinforcement Age at test

(days)

Compressive

strength (MPa)

Eccentricity,

ex (mm)

Eccentricity,

ey (mm)

1 4N12 94 37 15 25

2 4N12 403 45 15 50

3 4N12 432 47 30 70

4 8N12 446 59 35 35

5 8N12 453 53 50 40

6 8N12 404 58 70 50

7 4N12 87 50 15 25

8 4N12 367 52 15 50

9 4N12 411 48 30 70

10 8N12 418 63 35 35

11 8N12 446 62 50 40

12 8N12 397 61 70 50

Page 57: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

42

Figure 2.15 column test set-up (Rahman and Sarker 2011)

Although the geopolymer cylinders and columns were subjected to direct sun and

rain for more than a year, no change in the appearance of geopolymer cylinders and

columns was observed. This was due to the stability of geopolymer concrete as a

structural material in inconsistent atmospheric conditions. The geopolymer columns

performance in terms of load-deflection and failure subjected to biaxial bending was

identical to those observed for OPC concrete columns. The axial load capacity of

columns increased with the increase of concrete compressive strength and

reinforcement ratio and decrease of load eccentricity. The authors also reported using

Bresler’s reciprocal load formula to predict the failure load of the slender columns

subjected to bi-axial bending yielding conservative test-to-predicted ratio of 1.18.

Therefore, the analytical method gives conservative prediction of the capacity of

geopolymer concrete columns under bi-axial bending.

Page 58: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

43

The performance of geopolymer concrete columns with glass-fibre-reinforced-

polymer (GFRP) was examined by Maranan et al. (Maranan, Manalo et al. 2016). A

total of eight full-scale GFRP-reinforced geopolymer concrete columns were

fabricated including six short columns and 2 slender columns with different

transverse reinforcement layout as shown in Figures 2.16 and 2.17.

Figure 2.16 Details and configuration of the column specimens (Maranan, Manalo et al.

2016)

Figure 2.17 Configuration of the columns (Maranan, Manalo et al. 2016)

Page 59: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

44

The axial capacity of the geopolymer concrete column was higher than that of OPC

concrete column while negligible difference in terms of ductility and confinement

efficiency was observed. The spiral-confined columns showed a more ductile

behaviour and higher post-concrete-cover spalling strength compared to their

companion hoop-confined ones. The short column failed due to the crushing and/or

shears while slender columns collapsed at a load 66% and 82% of the strength of

their short-column companion due to the buckling (Figure 2.18).

Figure 2.18 Failure modes of columns (Maranan, Manalo et al. 2016)

Other geopolymer concrete members

Ganesan et al. (Ganesan, Indira et al. 2013) tested ten geopolymer concrete and ten

OPC concrete wall panels under uniformly distributed axial load in one-way in-plane

action to investigate the influence of slenderness ratio (SR) and aspect ratio (AR) of

the walls. The reinforcement arrangements of the panels are depicted in Figure 2.19.

Page 60: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

45

Figure 2.19 Reinforcement arrangement in the panel (Ganesan, Indira et al. 2013)

The geopolymer concrete wall panels failed in a more ductile pattern in comparison

with OPC samples as a result of presence of larger content of fine particles in their

matrix. The available models in ACI 318-08 for predicting the ultimate load of OPC

concrete wall panels gave conservative prediction in case of geopolymer concrete

wall panels.

Rajendran and Soundarapandian (Rajendran and Soundarapandian 2013) investigated

the flexural behaviour of geopolymer reinforced ferrocement slabs with varying

number of mesh layers and alkaline solution content. The cracking, yielding and

ultimate load of the geopolymer slabs were higher compared to the companion OPC

specimens. Furthermore, the geopolymer ferrocement slabs could sustain larger

deflection at yield and failure and also showed more micro-cracks and subsequently

exhibited ductility improvement (up to 26%) and energy absorption enhancement (up

to 109%) with respect to OPC slabs. In terms of cracking behaviour, the geopolymer

Page 61: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

46

specimens had increased number of cracks while the average crack width and

spacing were higher in OPC slabs.

To sum up, based on the published results, the existing guidelines and design

provisions could be used for design of geopolymer concrete members. Furthermore,

the general behaviour and failure of reinforced geopolymer members were identical

to those of OPC concrete specimens which further justify the use of available codes

of practice to design geopolymer concrete structural members.

Fire Resistance of Reinforced Concrete Members

Philosophy

Fire resistance rating requirements for buildings are specified in building codes such

as Building Code Requirements for Structural Concrete (ACI 318-14), National

Building Code of Canada (NRC, 1995), Eurocode 2 (2004), etc. Fire resistance can

be defined as the time for which an element of building construction is able to

withstand exposure to a standard temperature-time (ASTM E119 or ISO 834) and

load regime without a loss of its fire barrier function or load bearing function or both

(BS476, part 20, ACI 216).

The fire resistance of reinforced concrete members is generally developed by

employing prescriptive methods which are based on either standard fire resistance

tests or empirical calculation methods (Kodur, Dwaikat et al. 2009). The prescriptive

methods are less expensive than fire testing approaches but they are rigid and

restrictive and do not allow for engineering judgment. Furthermore, the current

prescriptive methods are limited to conventional concretes and may not be applicable

for structural members made of novel types of concrete such as geopolymer concrete.

Page 62: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

47

The fire safety in most countries is moving from prescriptive based method toward

performance based approach because it provides rational, innovative and cost

effective designs. However, the lack of calculation approaches limits the use of such

performance based fire safety design methods (Kodur, Dwaikat et al. 2009). There is

an increased focus on employing numerical methods mostly based on finite element

or finite difference methods to evaluate the fire performance of structural members

(Bratina, Čas et al. 2005, Kodur and Dwaikat 2008, Kodur, Dwaikat et al. 2009).

Spalling

The fire performance of concrete structural members is significantly influenced by

occurrence of fire-induced spalling (Dwaikat and Kodur 2009). The spalling might

happen immediately after exposure to rapid heating or might occur in further stages

while concrete has become so weak that cracks develop and concrete pieces come off

from surface (Kodur 2014).

The spalling might cause substantial reduction of strength and stiffness of reinforced

concrete members due to reducing the area of structural members and increase of

heat transfer to steel reinforcement (Kodur 2000). Based on the published data, many

researchers reported explosive spalling in concrete members under laboratory

conditions while few experiments exhibited little or no substantial spalling (Dwaikat

and Kodur 2009). Most researchers agree that spalling occurs mainly due to low

permeability of concrete and moisture migration at elevated temperatures (Kodur

2000).

Two major theories are available for explaining spalling phenomenon (Kodur 2000):

Pressure buildup: The buildup of pore pressure during heating causes the occurrence

of spalling. High strength concrete (HSC) is believed to be more susceptible to this

Page 63: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

48

pressure buildup than normal strength concrete (NSC) while this extremely high

water vapour pressure cannot escape through highly compact and low permeable

structure of HSC. Once this pore pressure exceeds the tensile strength of concrete,

breaking off of concrete layers might happen due to tensile fracture.

Restrained thermal dilatation: based on this mechanism, spalling happens due to the

restrained thermal dilatation close to heat surface which causes the compressive

stress development parallel to the heated surface and these stresses are then released

through spalling. The pore pressure could play a major role to initiate instability in

the form of explosive thermal spalling.

Based on the literature, the main reason for occurrence of spalling is buildup of pore

pressure and other mechanisms such as restrained thermal dilation might indirectly

influence the pore pressure buildup and temperature-dependent tensile strength of

concrete (Kodur 2014).

Fire resistance tests on reinforced OPC concrete columns

The first series of fire resistance experiments on forty-one large scale reinforced

concrete columns was conducted by Lie (Lie and Woolerton 1988). All columns

were 3810 mm long with different shape and dimensions including: square cross

section of 203, 305 and 406 mm, rectangular cross sections of 305 x 457 mm and

203 x 915 mm and circular cross sections of 355 mm diameter. The influence of

major parameters including: load level, percentage of longitudinal reinforcing steel,

effective length, concrete strength, relative humidity in concrete, column cross

sectional area, column shape (square, rectangular or circular), concrete aggregate

type, axial restraint of thermal expansion and load eccentricity, on the performance

of columns were investigated.

Page 64: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

49

The results showed that load level, cross sectional area for identical shapes and type

of aggregate have the most substantial effect on the fire performance of reinforced

concrete columns. The replacement of siliceous aggregate by carbonate aggregate

significantly increased the fire resistance of the columns. Furthermore, the

rectangular sections exhibited higher fire resistance properties compared to those of

square columns of same thickness. At the same load level, the change in the amount

of reinforcing steel and effective length of the columns marginally caused an

increase in fire resistance. Small eccentric load causes a small reduction in fire

resistance. The effects of concrete strength, concrete moisture content and restraint of

axial column expansion found to be insignificant.

Kodur and Mcgrath (Kodur and Mcgrath 2003) evaluated the fire performance of six

full-scale reinforced high strength concrete columns. All the square columns were of

3810 mm length, including three with cross section of 406 mm and the remaining

three were of 305 mm. It was reported that type of aggregate, concrete strength, load

intensity and ties configuration affect the fire resistance of HSC columns.

The tie layout (bending of ties at 135º ties and provision of cross ties) and closer

spacing significantly improved the fire resistance of HSC columns. The occurrence

of spalling was observed in the later stages of experiment.

In another study by these authors (Kodur and McGrath 2006), the effect of silica

fume and lateral confinement on fire resistance of HSC columns were evaluated. The

columns details including layout of ties are shown in Figure 2.20.

The results of this study highlighted the major influence of silica fume inclusion, ties

configuration and type of aggregate on the fire performance of HSC columns. The

columns containing higher amount of silica fume exhibited lower fire resistance and

Page 65: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

50

higher extent of spalling compared to columns with lower content of silica fume.

Moreover, by reducing the ties spacing and the provision of cross-ties and using the

carbonate type aggregate, increase of fire resistance and decrease of spalling was

observed.

Figure 2.20 Elevation and cross section of reinforced concrete columns (Kodur and

McGrath 2006)

The fire tests of circular concrete columns were conducted by Franssen and Dotreppe

(Franssen and Dotreppe 2003). A total of four circular columns with cross section of

300 mm and a length of 2100 mm were tested. Two of the columns were reinforced

with 6 Ф 20 longitudinal bars and two with 6 Ф 12. The surface spalling was

observed within the range of 20 and 60 minutes from the start of heating, regardless

of cross section shape and size of the longitudinal reinforcement. Despite the

occurrence of surface spalling, the achieved fire endurance values were relatively

Page 66: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

51

high. An increase of the load level substantially decreased the fire resistance of the

columns.

Xu and Wu (Xu and Wu 2009) examined the fire resistance of full-scale reinforced

concrete columns with different cross section shapes including four L-shaped, four

T-shaped, three +-shaped and one square shaped. The columns lay out are depicted in

Figure 2.21. The influence of axial load ratio and fire exposure condition on failure

mode, axial deformation and fire resistance of the columns were evaluated.

With the same load ratio, the fire resistance of columns ranged between 60% and

73% of that of the square column, so nonrectangular columns (NRCs) exhibited

lower fire resistance in comparison with rectangular columns. Furthermore, the axial

load ratio and fire exposure regime had substantial influence on the fire resistance of

reinforced concrete NRCs. By increasing the load ratio, a decrease in maximal

expansion deformation and the corresponding time to failure was achieved.

Page 67: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

52

Figure 2.21 Elevation and cross-section of the specimens (unit: mm) (Xu and Wu 2009)

Fire resistance of geopolymer concrete elements

Lyon et al. (Lyon, Balaguru et al. 1997) studied the fire performance of geopolymer

fibre composites. The maximum temperature capacity of carbon fibre reinforced

geopolymer composite was higher than 800 ℃ which proved to be much higher than

some other identical materials.

In terms of large scale experiments, Sarker and Mcbeath (Sarker and Mcbeath 2015)

conducted an experiment to evaluate the residual strength of reinforced fly ash based

geopolymer panels, subjected to standard ISO 834 fire on one side for two hours.

Totally six 500 mm square reinforced OPC and geopolymer concrete panels of 125,

150 and 175 mm thickness were fabricated and tested. The test set-up for

geopolymer panels is presented in Figure 2.22.

Page 68: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

53

Figure 2.22 A concrete test panel set for fire exposure (Sarker and Mcbeath 2015).

It was found that the heat transfer at high temperature was faster in geopolymer

concrete panel than its OPC counterpart which led to smaller temperature differential

in the geopolymer concrete panel than OPC ones. Also the damages caused by

cracking and spalling were less in the geopolymer panels as shown in Figures 2.23

and 2.24.

Page 69: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

54

Figure 2.23 Failure of the 150 mm OPC concrete panel (Sarker and Mcbeath 2015).

Figure 2.24 Failure of the 150 mm geopolymer concrete panel (Sarker and Mcbeath 2015).

Page 70: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

55

The results of residual compression tests after cooling down for the panels indicated

that the geopolymer panels retained higher percentage (66%) of their strength than

OPC concrete panels (52%). The higher residual strength of reinforced geopolymer

panels was associated with the less internal damage due to the less temperature

differential than in the OPC concrete specimens.

Concluding Remarks

From the literature review it was found that geopolymer concrete has many

advantages that make it a practical alternative to OPC concrete. The results of the

experiments on the behaviour of geopolymer concrete members such as beams and

columns showed that the design provisions for OPC concrete members could be used

in case of geopolymer concrete members as in most cases conservative prediction of

ultimate load-capacity of geopolymer concrete members was observed. Furthermore,

the overall performance of reinforced geopolymer members under loading were

identical to those of OPC concrete specimens which further promote the use of

existing codes of practice to design geopolymer concrete structural members.

The development of one-part geopolymer concrete, as a new class of geopolymers,

with similar production to that of OPC concrete, encourages the widespread

application of geopolymers in construction industry. Therefore, further research is

required to evaluate the structural performance of one-part geopolymer concrete

members.

Moreover, very limited reports are available on the fire tests and finite element

models that evaluate the behaviour of geopolymer concrete materials and structural

members in fire as one of the major concerns in high-rise building and infrastructure

design. It is necessary to evaluate the extent of damage of steel reinforced one-part

Page 71: Mechanical, Thermal and Structural Investigations of One

Chapter 2 Literature Review

56

geopolymer concrete elements at high temperatures since real structures are mostly

made of reinforced concrete members. Therefore, further research efforts should be

undertaken to clarify the potential improvement in fire resistance of geopolymer

concrete members.

Page 72: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

57

Chapter 3 Development of Hybrid OPC-Geopolymer

Concrete

General

One-part hybrid ordinary Portland cement (OPC)-geopolymer concrete mixes were

developed as part of this research and reported in this chapter, in which solid

potassium carbonate (7.5 wt.% of the total geopolymeric raw materials) was used as

the main activator and the OPC as a source of silicate and poly silicate was blended

with geopolymeric raw materials (fly ash and ground granulated blast furnace slag)

in different proportions. The influence of OPC content on the workability, setting

time, compressive strength development and microstructure of the concrete mixes

was investigated.

This chapter includes the introduction of materials, specimen preparation, testing

procedures, test results and discussion, followed by the conclusion of the chapter.

Experimental Procedure

Materials

The low calcium fly ash (ASTM C 618 Class F) used in this research was sourced

from Gladstone Power Station, Queensland, Australia. Commercially available slag

supplied by Australian Builders and OPC (ASTM C 150 Type 1) supplied by Cement

Australia were also used in the experiments. The results of X-ray Fluorescence

(XRF) analyses of the raw materials are presented in Table 3.1. Natural river sand

with a specific gravity of 2.6 and water absorption of 3.5% was used as fine

aggregate, whereas crushed basalt with a maximum nominal size of 10 mm, a

Page 73: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

58

specific gravity of 2.8 and water absorption of 1.6% was used as coarse aggregate.

Prior to mixing, all aggregates were dried in an oven at 105℃ for a period of 48

hours to remove any moisture.

Table 3.1 Chemical composition of OPC, fly ash and slag determined by XRF

Material SiO2 Al2O3 Fe2O3 CaO MgO Na2O K2O SO3 TiO2 LoIa

OPC (%) 20.2 4.9 2.6 62.7 2.0 0.2 0.4 2.2 − 2.0

Fly ash (%) 50.3 27.56 12.68 2.7 1.3 0.6 0.7 0.36 1.42 0.67

Slag (%) 32.5 13.0 0.55 43.2 4.7 0.2 0.3 4.3 0.51 0.08

a Loss of ignition

This study has used three different types of solid activators, such as potassium

carbonate, calcium hydroxide and sodium silicate. Powder form potassium carbonate

denoted as “K2CO3” with 99.99% purity was supplied by Sigma-Aldrich in Australia.

An industrial grade of calcium hydroxide Ca(OH)2 commonly used in the

construction industry was supplied by Cement Australia. The sodium silicate-based

activator denoted as “Na2SiO3-Anhydrous” was supplied by Redox Australia. The

supplied sodium silicate has the chemical composition: 51% Na2O, 46% SiO2 and

3% H2O with a modulus (Ms = SiO2/Na2O) of 0.9.

Concrete mix proportions and specimen preparation

The proportions of one-part hybrid OPC-geopolymer concrete mixes together with

their control mixes (with no activator) are illustrated in Table 3.2. All the hybrid

OPC-geopolymer mixes included solid activator of potassium carbonate with a

weight of 7.5% of the total geopolymeric raw materials; this amount was determined

based on several trial tests. The hybrid OPC-geopolymer mixes were designated with

the percentage of OPC in the binder. Accordingly, for mixes GP60, GP40, GP30,

GP20, GP10 and GP0, the amounts of OPC were 60 wt.%, 40 wt.%, 30 wt.%, 20

Page 74: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

59

wt.%, 10 wt.% and 0% of the total binder, respectively. For a corresponding control

mix, “GP” in the designation is simply replaced by “C”. For example, C60 without

activator is the control mix of GP60. It should be noted that the replacement of fly

ash and slag by OPC was based on the weight of the materials and the difference in

material densities was not considered in this study. It is estimated that the material

replacement caused a variation in the total mixture volume up to 0.9%, which is

negligible. This simplification was also adopted in (Deb, Nath et al. 2014, Noushini,

Aslani et al. 2016) to adjust their mix design. Furthermore, three additional concrete

mixes A1−3 were prepared to evaluate the influence of combining calcium hydroxide

or sodium silicate with potassium carbonate as activator on the strength

development. The OPC percentage was kept constant at 10% by weight of the total

binder in these concrete mixes. Apart from the use of potassium carbonate (27

kg/m3), additional sodium silicate and calcium hydroxide at a weight of 2.5 wt.%, 5

wt.% and 2.5 wt.% of the total geopolymeric raw materials were added to mixes A1,

A2 and A3, respectively. The amount of sodium silicate in mix A2 was 18 kg/m3,

which is twice that in mix A1.

All the concrete mixes in this study had the same amount of binder and aggregate

and a constant water to binder (w/b) ratio of 0.3. The aggregates proportions shown

in Table 3.2 were based on the saturated surface dry (SSD) condition. Furthermore,

the slag content of all mixes was kept constant at 10 wt.% of the total geopolymeric

raw materials, although future research might be required to investigate the variation

of slag content as well.

All the mixes were prepared using a vertical pan mixer. Solid ingredients, including

the OPC, fly ash, slag, solid activator, citric acid, coarse aggregate and sand were

Page 75: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

60

added to the mixer and dry-mixed thoroughly for 3 min before adding water. The

water and superplasticiser were then added gradually and mixing was continued for

another 4-6 min until a uniform mix was achieved (Figure 3.1). The freshly mixed

concrete was then cast in 100 × 200 mm cylindrical moulds in two layers and each

layer was compacted on a vibrating table to achieve proper consolidation (Figure

3.2). The samples were demoulded 24 hours after casting and were wrapped with

plastic sheet. Then, they were stored in a controlled room with temperature of

20−23 ℃ and relative humidity of 65±10% until testing.

In addition to the concrete mixes in Table 3.2, one-part hybrid OPC-geopolymer

pastes were also prepared with the same proportions of binder and activator to

measure the initial and final setting times and to be used for microstructural analysis.

The pastes were mixed manually in a laboratory bowl until a uniform mixture was

achieved.

Figure 3.1 Mixing one-part hybrid OPC-geopolymer concrete

Page 76: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

61

Table 3.2 Mix proportions of one-part geopolymer concrete (kg/m3).

Mix

No.

Designation Coarse

aggregate

Sand OPC Fly

ash

Slag Activator Citric

acid

Water Superplasticiser

1 GP60a 1,220 635 240 144 16 12 0.96 120 6

2 C60 1,220 635 240 144 16 - 0.96 120 6

3 GP40a 1,220 635 160 216 24 18 0.64 120 6

4 C40 1,220 635 160 216 24 - 0.64 120 6

5 GP30a 1,220 635 120 252 28 21 0.48 120 6

6 C30 1,220 635 120 252 28 - 0.48 120 6

7 GP20a 1,220 635 80 288 32 24 − 120 6

8 C20 1,220 635 80 288 32 - - 120 6

9 GP10a 1,220 635 40 324 36 27 − 120 6

10 C10 1,220 635 40 324 36 - - 120 6

11 GP0a 1,220 635 − 360 40 30 − 120 6

12 A1b 1,220 635 40 324 36 36 − 120 6

13 A2b 1,220 635 40 324 36 45 − 120 6

14 A3c 1,220 635 40 324 36 36 − 120 6

a Potassium carbonate (powder form) was used as activator

b Potassium carbonate and sodium silicate (powder form) were combinedly used as activator

cPotassium carbonate and calcium hydroxide (powder form) were combinedly used as

activator

Figure 3.2 Concrete cylinders cast in steel cylindrical moulds

Page 77: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

62

Testing procedure

The workability of fresh concrete mixes was tested by slump test immediately after

mixing, according to AS 1012.3.1 (1998) procedure (Figure 3.3). The setting time of

the paste was measured at a temperature of 21−23℃ using a Vicat apparatus in

accordance with ASTM C 191 (2008) as depicted in Figure 3.4. Compressive

strength tests were conducted at 3, 7 and 28 days of age using a universal testing

machine as shown in Figure 3.5. For each age, three cylindrical samples were tested

and the reported results are the average of the three. Cylindrical specimens were

tested at a loading rate of 20±2 MPa/min according to AS1012.9 (1999). All

cylinders were ground flat on both ends to ensure having smooth and parallel top and

bottom faces prior to performing the tests.

Figure 3.3 Conducting slump test for one-part hybrid OPC-geopolymer concrete mixes

Page 78: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

63

Figure 3.4 Vicat apparatus for measuring setting time of paste

Figure 3.5 Test setup for compressive strength test

Page 79: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

64

To observe the microstructure of the hybrid OPC-geopolymer paste samples, a JEOL

6510LV scanning electron microscope (SEM) was utilised. The elemental

composition of samples was observed using a Moran Scientific microanalysis EDS

(energy dispersive spectroscopy) system. XRD patterns were achieved implying a

Bruker D8 Advance Power Diffractometer. Diffraction analysis were made from 15

to 60 2𝜃 using copper K radiation. The excitation voltage was 40 kV at 40 mA,

counting was for 5 seconds.

Results and Discussion

Behaviour of fresh mixes

The slump test values for one-part hybrid OPC-geopolymer concrete mixes are

presented in Figure 3.6. It can be observed that OPC content had a significant effect

on the workability of the one-part mixes. The geopolymer concrete mix (GP0) that

contains no OPC exhibited the highest workability as indicated by the slump value of

122 mm. A decrease in slump was observed with increasing OPC content when the

ratio of activator to geopolymeric raw materials was kept at 7.5%. The decrease was

16% and 22% for mixes with 10 wt.% and 20 wt.% OPC addition compared to GP0

(without OPC addition). The slump dropped significantly for GP30 with 30 wt.%

OPC. A very low slump of 30 mm was observed for mix GP60 with 60 wt.% OPC.

The decrease in slump value with increasing OPC could be due to the rapid reaction

of OPC with the alkaline activator. Addition of superplasticiser might be helpful to

improve the workability of hybrid mixes with high OPC content.

Page 80: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

65

Figure 3.6 Effect of OPC inclusion on the slump

Initial and final setting times were measured for the corresponding pastes of mixes

GP60, GP40, GP30, GP20, GP10, and GP0, and the obtained results are shown in

Figure 3.7. For convenience of description, the mix designations presented in Table

3.2 are also used for paste samples in this section. Mix GP0 which contained only fly

ash and slag as binder did not set within the first 24 hours as a result of slow rate of

chemical reaction at ambient temperature. By incorporating OPC in the mixes, both

the initial and final setting times were significantly decreased which is in agreement

with previous studies (Suwan and Fan 2014, Nath and Sarker 2015). It has been

reported that the initial setting times were only 22 and 11 min when the OPC to the

total binder ratios were increased to 50% and 70%, respectively (Suwan and Fan

2014). The quick setting of the OPC-rich mixes would prevent transportation and

proper casting of the concrete. In this study, different amounts of citric acid were

added to mixes GP60, GP40 and GP30. The OPC hydration in these mixes was

effectively retarded and the setting time was increased because of the use of citric

acid. For mixes with 60% (GP60), 40% (GP40) and 30% (GP30) of OPC in the

binder, the initial setting times are 30, 39, and 44 min and the final setting times are

0

20

40

60

80

100

120

140

GP0 GP10 GP20 GP30 GP40 GP60

Slu

mp

(m

m)

Page 81: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

66

53, 75, and 107 min, respectively. In contrast, the mixes with a lower content of OPC

generally required longer setting times although no citric acid was used. The initial

setting times are 61 and 86 min and the final setting times are 102 and 139 min for

mixes with 20% (GP20) and 10% (GP10) of OPC in the binder, respectively. It can

be concluded that the incorporation of OPC in the mixture can accelerate the setting

of one-part geopolymer concrete. If the OPC content in the binder is 30% or above, a

suitable amount of citric acid can be added to retard the OPC hydration.

Figure 3.7 Setting times of one-part pastes.

Compressive strength

The measured compressive strengths at 1, 7 and 28 days are presented in Table 3.3

for all hybrid OPC-geopolymer concrete and their control mixes. Mix GP0 without

OPC was found to be very weak and only achieved a low strength of 11.4 MPa at 28

days. In contrast, the increase in OPC percentage significantly improved the

corresponding compressive strength at various ages for both hybrid OPC-geopolymer

and control mixes. In general, the compressive strengths of the hybrid OPC-

geopolymer concrete mixes are higher than those of the control mixes without

0

20

40

60

80

100

120

140

160

GP60 GP40 GP30 GP20 GP10 GP0

Sett

ing

tim

e (

Min

)

Initial

Final

> 2

4 h

Page 82: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

67

activator regardless of their OPC content. This strength enhancement is even more

significant at early ages of 1 day and 7 days. At 28 days, mixes GP10, GP20, GP30,

GP40 and GP60 achieved compressive strengths of 33.4, 39.7, 50.3, 52.5 and 55.0

MPa, respectively. In contrast, the control mixes C10, C20, C30, C40 and C60

achieved lower compressive strengths of 18.3, 24.3, 33.0, 38.8 and 44.2 MPa,

respectively. This comparison confirms the contribution of alkali activation to

strength development of the hybrid OPC-geopolymer concrete. Meanwhile, it is

found that the percentage of strength increase at 28 days due to alkali activation

decreased from 82.5% to 24.4% as the OPC content increased from 10% to 60%.

The quick hardening of the geopolymer mixes in the presence of calcium compounds

with respect to the control mixes could be attributed to the rapid reaction of OPC

with the alkaline activator, which consumes the expelled water from the

geopolymerisation process and consequently promotes this process (Nath and Sarker

2015, Assi, Ghahari et al. 2016). The hydration heat of OPC might also accelerate

the alkali activation, leading to increased strength development. Clearly, when the

OPC content in the binder reaches 10% or higher, the 28-day compressive strength of

the corresponding concrete exceeds 32 MPa, ensuring the structural use of this type

of concrete.

Page 83: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

68

Table 3.3 Compressive strengths of geopolymer concrete mixes at 1, 7 and 28 days

Mix No. Designation 1 day 7 days 28 days

1 GP60 27.3±0.1 45.3±0.3 55.0±1.8

2 C60 7.3±0.9 28.4±1.4 44.2±1.5

3 GP40 23.8±1.4 44±2.6 52.5±0.4

4 C40 6.6±0.8 25.6±0.2 38.8±1.2

5 GP30 25.3±0.3 45.6±0.5 50.3±1.3

6 C30 5.8±1.0 24.2±0.5 33. 0 ±0.6

7 GP20 17.8±1.1 31.3±0.8 39.7±0.8

8 C20 4.1±1.3 17.8±1.2 24.3±0.6

9 GP10 13.2±1.9 26.7±0.4 33.4±0.6

10 C10 3.2±0.4 12.8±1.2 18.3±2.8

11 GP0 0±0 2.3±0.7 11.4±0.9

The current research also indicates that, when the OPC content in the binder reaches

10% or higher, the compressive strength of a hybrid OPC-geopolymer mix is

comparable to that of conventional geopolymer using alkaline solution of sodium

silicate, sodium hydroxide or their combination (Part, Ramli et al. 2015, Singh,

Ishwarya et al. 2015). For example, mix GP30 with 30% of OPC in the binder

achieved a compressive strength of 25.3 MPa at 1 day and 50.3 MPa at 28 days. It is

worth noting that this mix only used potassium carbonate powder as activator at 7.5

wt.% of the total geopolymeric raw materials. In contrast, the corresponding

percentage in conventional geopolymer normally ranges from 13 wt.% to 16 wt.%

(Rangan 2008). As can be seen, the current one-part hybrid OPC-geopolymer

concrete used much less activator compared with the conventional geopolymer.

Page 84: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

69

Therefore, the developed hybrid mixture is more feasible and economical for on-site

applications.

The influence of additional calcium hydroxide and sodium silicate in the activator on

the strength development is demonstrated in Figure 3.8. Compared with mix GP10,

the addition of sodium silicate in mixes A1 or A2 has only moderate influence on the

concrete strength. Although the amount of sodium silicate was doubled in A2 as

compared to A1, the strength increase of 4.2% at 28 days was marginal. It seems that

sodium silicate has no obvious influence on the alkali reaction in the hybrid OPC-

geopolymer concrete. However, the strength increase is more significant when

additional calcium hydroxide (2.5 wt.% of the total geopolymeric raw materials) was

added to mix A3, which is attributable to the increased pH value of the solution to

promote geopolymerisation. Compared with the 28-day strength of GP10 (33.4

MPa), the corresponding strength of mix A3 was increased to 37.6 MPa.

Figure 3.8 Influence of calcium hydroxide and sodium silicate on strength development

0

5

10

15

20

25

30

35

40

0 5 10 15 20 25 30

Co

mp

ress

ive

stre

ngt

h (

MP

a)

Age (day)

A1

A2

A3

GP10

Page 85: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

70

Microstructural analysis

The effect of OPC content on the microstructure of the one-part mixes was studied.

Figure 3.9 presents the microstructural images of the samples at 28 days of age using

backscattered electron (BSE). As can be seen in the SEM micrographs, increasing

the OPC content in the mix leads to a more compact and less porous structure.

Meanwhile, there are less unreacted spherical fly ash particles in the matrix. Because

of the inclusion of a large amount of OPC, mix GP60 shown in Figure 3.9 also

comprises both unreacted and partially reacted OPC particles as confirmed by the

EDS results. Komnitsas (Komnitsas and Zaharaki 2007) also reported that, when the

OPC concentration is high, microstructure porosity decreases for the formation of

amorphous Ca-Al-Si gels, leading to increased strength. It can also be seen in Figure

3.9 that unreacted or partially reacted OPC particles are rarely seen in the paste

having 30% (GP30) or lower amount of OPC, which shows their participation in the

formation of either geopolymer gel or a new gel. The microstructure of ambient

cured GP0 contains both unreacted fly ash particles and geopolymeric matrix,

demonstrating a loose and amorphous structure. The large amount of unreacted fly

ash particles explains the very low compressive strength as confirmed by the

compression tests. GP30 has a more compact microstructure with less unreacted fly

ash particles with respect to GP0, GP10 and GP20, although it has a more porous

microstructure than GP60 and GP40. For GP30, the major product of this low

calcium content structure is geopolymeric gel surrounded by partial CSH gel in the

entire network as was also observed previously (Suwan and Fan 2014).

Overall, the SEM images depict the coexistence of geopolymeric gel and CSH gel

phases, showing a hybrid product like calcium alumino-silicate hydrate (CASH) as

Page 86: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

71

suggested by Garcia-Lodeiro et al. (Garcia-Lodeiro, Palomo et al. 2011). This will be

further confirmed from the C/S ratios measured as follows.

Figure 3.9 BSE images of hybrid OPC-geopolymer mixes

Quantitative EDS analysis was performed on the hybrid OPC-geopolymer paste

samples at 28 days to evaluate the reaction products of mixes. The oxide molar ratios

presented in Table 3.4 are the averages measured at 5 different spots. The EDS

Page 87: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

72

analysis indicates that the CaO-to-SiO2 (C/S) ratio decreases with decreasing OPC

content in the hybrid OPC-geopolymer mixes. The C/S ratio is 1.1 in GP60 and

decreases to 0.23 for GP10, whereas the corresponding ratio of GP0 is only 0.06. By

adding calcium hydroxide into the mix, the C/S ratio slightly increases from 0.23 in

GP10 to 0.29 in A3. Obviously, a mix with a high C/S ratio has a high potential for

the formation of CSH, leading to increased early strength. For neat OPC paste, the

C/S ratio was reported to vary in the range of 1.2-2.3 (Yip and Van Deventer 2003),

which is higher than those found in this study (0.18-1.1). It can be inferred that the

nature of CSH formed in the hybrid mixes is different from that of normal CSH

formed in OPC.

Table 3.4 Oxide molar ratios in different mixes

Mix

No.

Designation SiO2/Al2O3 CaO/SiO2 CaO/Al2O3 K2O/SiO2 K2O/Al2O3

1 GP60 2.51 1.10 2.78 0.16 0.41

3 GP40 2.50 0.72 1.81 0.20 0.50

5 GP30 2.47 0.53 1.32 0.22 0.54

7 GP20 2.41 0.37 0.89 0.23 0.55

9 GP10 2.39 0.23 0.56 0.23 0.55

10 GP0 2.40 0.06 0.14 0.24 0.57

12 A1 2.52 0.19 0.46 0.23 0.57

13 A2 2.53 0.18 0.47 0.24 0.60

14 A3 2.47 0.29 0.72 0.23 0.56

The XRD graphs of hybrid OPC-geopolymer pastes are plotted in Figure 3.10. In

general, the hydrated OPC contains major crystalline phases of CSH while

geopolymer concrete is a mixture of crystalline phases of Quartz and Mullite and

amorphous phases observed in the region of 20 − 35° 2𝜃. Previous studies showed

Page 88: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

73

that CSH and NASH (main reaction product of class F fly ash activation) gels are

compatible and precipitation of a mix of gels is responsible for the setting and

hardening of this type of hybrid concrete (Palomo, Fernández-Jiménez et al. 2007).

The OH- concentration and the mix composition were reported as controlling factors

to form CSH/semi–crystalline phase or NASH and CASH (main product of fly ash

activation reaction)(Palomo, Fernández-Jiménez et al. 2007). Peaks due to the

crystalline components of Quartz and Mullite from the fly ash were observed in all

the pastes and the intensity of their peaks decreases with increasing OPC content in

the mix. On the other hand, CSH, CASH, Nepheline, Hatrurite, Calcite and

Portlandite appear and their intensity consistently increase with increasing OPC

content. In all mixes, a broad and amorphous hump was observed, indicating the

coexistence of geopolymeric and hydrated OPC products in the binder system. The

XRD results for mixes GP30, GP20 and GP10 differ from the results achieved for

mixes GP60 and GP40 as Portlandite does not appear as a hydration product in

GP30, GP20 and GP10. In general, it can be concluded that the one-part hybrid OPC-

geopolymer mixes had different reaction products with respect to OPC and

conventional geopolymers. This is consistent with previous research findings (Suwan

and Fan 2014, Nath and Sarker 2015). The initial reaction of calcium compounds

with activator resulted in the quick formation of both amorphous CSH and semi-

crystalline geopolymeric network. In the later stage, the CSH gel increasingly formed

due to the decrease of alkalinity in the mix. The results of SEM, EDS and XRD

confirm that the hybrid OPC-geopolymer mix produces a mixture of cross-linked

network of geopolymeric and CSH gels in the binder system.

Page 89: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

74

Figure 3.10 XRD analysis of various hybrid OPC-geopolymer systems at the age of 28 days.

Conclusions

This study aimed to develop ambient temperature cured one-part hybrid OPC-

geopolymer concrete mixes using ordinary Portland cement (10−60 wt.% of the total

binder) as initial binder in combination with solid potassium carbonate (7.5 wt.% of

the total geopolymetric raw materials) as alkaline activator. The following

conclusions can be drawn within the scope of this study:

1. Increasing OPC content decreased workability of one-part geopolymer concrete

mixes. The geopolymer concrete mix (GP0) obtained the highest slump (122 mm)

while the mix containing 60% OPC (GP60) achieved the lowest value of 30 mm. The

decrease in slump value with increasing OPC could be due to the rapid reaction of

OPC with the alkaline activator.

Page 90: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

75

2. By incorporating OPC in the mixes with potassium carbonate, both the initial and

final setting times were significantly decreased as the presence of OPC accelerated

the geopolymerisation reaction. In contrast, the mix containing only fly ash and slag

as binder did not set within the first 24 h as a result of slow rate of chemical reaction

at ambient temperature. If the OPC content in the binder is 30% or above, a suitable

amount of citric acid can be added to retard the OPC hydration.

3. The compressive strengths of the hybrid OPC-geopolymer concrete mixes are

higher than those of the control mixes without activator regardless of their OPC

content. This strength enhancement is even more significant at early ages of 1 day

and 7 days. The percentage of strength increase at 28 days due to alkali activation

decreased from 82.5% to 24.4% as the OPC content increased from 10% to 60%.

When the OPC content in the binder reaches 10% or higher, the 28-day compressive

strength of the corresponding concrete exceeds 32 MPa, ensuring the structural use

of this type of concrete.

4. The type of activator has some influence on the concrete strength. An increase

was observed when additional calcium hydroxide (2.5 wt.% of the total

geopolymeric raw materials) was used. However, while additional sodium silicate

(2.5 wt.% and 5 wt.% of the total geopolymeric raw materials) was added, the

increase in compressive strength was moderate. Further research is needed to

determine the optimised amount and combinations of activators to be used in the mix

of the one-part hybrid OPC-geopolymer concrete.

5. Microstructural investigation showed that hybrid OPC-geopolymer mixes had

different final product compared to OPC and geopolymers. The results of SEM, EDS

and XRD confirm that the hybrid OPC-geopolymer mix produces a mixture of cross-

Page 91: Mechanical, Thermal and Structural Investigations of One

Chapter 3 Development of Hybrid OPC-Geopolymer Concrete

76

linked network of geopolymeric and CSH gels in the binder system, which allows the

concrete to develop high early and ultimate strength.

Overall, the developed one-part hybrid OPC-geopolymer concrete could be a suitable

material for on-site operation as the need for heat curing and the use of highly

alkaline solutions have been eliminated. The setting time, workability, compressive

strength and microstructure properties can be controlled by adjusting the OPC

content and activator concentrations. Further investigations could be conducted to

study other properties of the hybrid OPC-geopolymer concrete, such as shrinkage,

creep, thermal properties and durability, etc.

Page 92: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

77

Chapter 4 Development of One-Part Geopolymers

Introduction

One-part geopolymer mixes were developed as part of this research and reported in

this chapter, in which highly corrosive alkali solutions used as activators in

conventional geopolymers were replaced by different solid activators (in powder

form), such as sodium silicate, calcium hydroxide, sodium oxide, lithium hydroxide,

potassium carbonate or their combinations and blended with geopolymeric raw

materials (fly ash and ground granulated blast furnace slag) in different proportions.

The influence of slag content and type/combination of solid activators on the on the

workability, early and final strength, hardened density and microstructure of these

one-part geopolymer mixes were examined.

This chapter includes the introduction of materials, specimen preparation, testing

procedures, test results and discussion, followed by the conclusion of the chapter.

Theoretical Background

In this experiment, solid alkaline activators such as sodium silicate, calcium

hydroxide, sodium oxide, lithium hydroxide and potassium carbonate as well as their

combinations were used to develop one-part geopolymer mixes. Sodium silicate and

calcium hydroxide have been extensively used as alkaline activators in previous

studies . Solid sodium oxide (Na2O) is more reactive than sodium hydroxide and its

hydration is exothermic, which may in return accelerate curing of geopolymer at

ambient temperature. The dissolution reaction of Na2O in water is shown in Equation

4.1:

Page 93: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

78

Na2O + H2O → 2 NaOH (4.1)

Sodium hydroxide (NaOH) is a highly alkaline activator for developing geopolymer

systems, but is mostly used in solution form because of its high hygroscopic

properties (Rattanasak and Chindaprasirt 2009, Somna, Jaturapitakkul et al. 2011).

The use of potassium salts, such as potassium hydroxide (KOH) and potassium

silicate (K2SiO3), in geopolymer systems has been found to show a greater level of

alkalinity compared to NaOH (Singh, Ishwarya et al. 2015). In this study, solid

potassium carbonate was chosen as one type of solid activator as its effectiveness has

been proven in our previous effort to develop hybrid OPC-geopolymer concrete

(Askarian, Tao et al. 2018). Furthermore, the use of lithium compounds has been of

great interest to researchers due to its outstanding characteristics such as high

alkalinity and the ability to mitigate the alkali-silica reaction (ASR) expansion

problem (McCoy and Caldwell 1951, Mo, Yu et al. 2003, Feng, Thomas et al. 2005,

Taha and Nounu 2008). It has been demonstrated that lithium could reduce the

dissolution of reactive silica by coating the aggregate particles and reduce the

opportunities for dissolved reactive silica to form ASR gel (Taha and Nounu 2008).

The feasibility of using lithium based activators in production of geopolymers has

been investigated previously and their effectiveness has been confirmed (O’Connor

and MacKenzie 2010, Chen, Li et al. 2012). Moreover, one-part alkali-activated

materials have been synthesised with lithium aluminate and laser-based additive

manufacturing (3D printing). Compared to one-part mixes with sodium aluminate,

lithium aluminate mixes showed lower solubility and faster hardening at room

temperature. However, shrinkage was an issue occurring with lithium aluminate

mixes (Sturm). In this study, the inclusion of non-hygroscopic lithium hydroxide

reactant in powder form in combination with other solid activators was also

Page 94: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

79

evaluated. It is expected that the ASR effect in geopolymer concrete made with high

slag content could be potentially mitigated by the inclusion of lithium hydroxide in

the mix (Singh and Ishwarya 2019).

Experimental Procedure

Materials

The low calcium fly ash (ASTM C 618 Class F) used in this research was sourced

from Gladstone Power Station, Queensland, Australia. Commercially available slag

supplied by Australian Builders and OPC (ASTM C 150 Type 1) supplied by Cement

Australia were also used in the experiments. The results of X-ray Fluorescence

(XRF) analyses of the raw materials are presented in Table 4.1. Natural river sand

with a specific gravity of 2.6 and water absorption of 3.5% was used as fine

aggregate, whereas crushed basalt with a maximum nominal size of 10 mm, a

specific gravity of 2.8 and water absorption of 1.6% was used as coarse aggregate.

Prior to mixing, all aggregates were dried in an oven at 105℃ for a period of 48

hours to remove any moisture.

Table 4.1 Chemical composition of OPC, fly ash and slag determined by XRF

Material SiO2 Al2O3 Fe2O3 CaO MgO Na2O K2O SO3 TiO2 LoIa

OPC (%) 20.2 4.9 2.6 62.7 2.0 0.2 0.4 2.2 − 2.0

Fly ash (%) 50.3 27.56 12.68 2.7 1.3 0.6 0.7 0.36 1.42 0.67

Slag (%) 32.5 13.0 0.55 43.2 4.7 0.2 0.3 4.3 0.51 0.08

a Loss of ignition

The sodium silicate (Na2SiO3 – GD) was supplied by PQ Australia with the chemical

composition of: Na2O = 27%, SiO2 = 54% and H2O = 19% with a modulus ratio (Ms

= SiO2/Na2O) of 2.0. Previous studies (Liew, Heah et al. 2017) highlighted the

Page 95: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

80

substantial influence of sodium silicate modulus on the pH, viscosity and setting time

of geopolymer mixes; a higher modulus leads to increase in the concentration of

three dimensional siloxane networks in the geopolymerisation process. Therefore,

Na2SiO3 – GD with a Ms of 2.0 was chosen in this study rather than other sodium

silicate products with low modulus ratios such as Na2SiO3 – Anhydrous (Ms = 0.9)

and Na2SiO3 – Penta (Ms=1.0). An industrial grade of calcium hydroxide Ca(OH)2

was supplied by Orica Chemicals in Australia. Fine powder of potassium carbonate

denoted as “K2CO3” with 99.99% purity, regent grade lithium hydroxide denoted as

“LiOH” with 98% purity and sodium oxide denoted as “Na2O” were provided by

Sigma-Aldrich, Australia.

Mix proportions and mixing

The proportions of the one-part geopolymer mixes are illustrated in Table 4.2. For all

the mixes, the amount of sodium silicate activator was set at 12 wt% of the binder

(fly ash or fly ash-slag blend) which was determined based on our trial tests and

previous findings in this area (Nematollahi, Sanjayan et al. 2015). In mixes M1 and

M9, sodium silicate (in powder form) was utilised as the sole alkaline activator,

whereas sodium silicate was combined with other solid activators in the rest mixes.

In mixes M1 to M8, only fly ash was used as the geopolymer precursor, whilst 50

wt% of fly ash in mixes M9 to M11 was replaced by slag to evaluate the influence of

precursor material on the properties of geopolymer. Therefore, M1 serves as the

reference mix, which consisted of only sodium silicate as activator and fly ash as

geopolymer precursor. In contrast, mixes M2 and M3 included additional 12 wt%

and 9 wt% Ca(OH)2, respectively, based on the weight of the binder. Because the

addition of Ca(OH)2 would reduce the workability significantly, the water/binder

(w/b) ratio of M2 and M3 was increased to 0.38 compared with 0.20 for M1. Mix M4

Page 96: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

81

had similar proportions to mix M3, except for the addition of 0.6 wt% lithium

hydroxide as activator. The difference between mixes M5 and M4 is that Ca(OH)2 in

M4 was replaced by the same amount of Na2O in M5. Half the Na2O in M5 was

further replaced by the same amount of Ca(OH)2 in M6. For mixes M7 and M8, the

amounts of sodium silicate and calcium hydroxide were the same as those in mix

M3. However, additional potassium carbonate was introduced as activator at 6 wt%

and 10 wt% of the binder for M7 and M8, respectively. The reference mixes for M9,

M10 and M11 are M1, M3 and M4, respectively. By comparing mixes M9−M11

with their reference mixes, the effect of substitution of fly ash with slag can be

investigated. It should be noted that adding slag could significantly reduce the

workability, so the w/b ratios of M9−M11 were increased compared to the

corresponding mixes with only fly ash.

All the mixes were prepared using a Hobart mixer. Solid ingredients, including fly

ash, slag and solid activator were added to the mixer and dry-mixed thoroughly for 2

minutes before adding water. The water and superplasticiser (if any) were then added

gradually and mixing was continued for another 6-10 minutes until a uniform mix

was achieved (Figure 4.1). It should be mentioned that due to the presence of sodium

silicate, extra time was required for mixing to achieve homogenous mixes.

Flowability of fresh one-part geopolymer pastes was measured through conducting

mini slump tests according to ASTM C1437 (1999). The top and bottom diameters

and the height of the conical mould used in this study were 70, 100 and 50 mm,

respectively. The fresh paste was poured in two layers into the mould placed on an

even steel plate. A tamping rod was used to tamp each layer 20 times to assure the

compaction, and the top surface was then levelled to remove any extra paste. After

Page 97: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

82

about 1 min, the mould was lifted and the flow diameter was determined as the mean

of measurements in two perpendicular directions. The relative slump value was

determined according to Equation 4.2:

𝛤𝑝 = (𝑑

𝑑°)2 − 1 (4.2)

where 𝛤𝑝 is the relative slump, d is the average of two measured diameters and 𝑑° is

the bottom diameter of the conical mould (100 mm in this study) (Nematollahi and

Sanjayan 2014).

After the measurement of flowability, the paste mixtures were cast in 50 × 100 mm

cylindrical moulds and vibrated to release any residual air bubbles. Then the samples

were covered with polyethylene sheets and divided into two groups subjected to two

different curing regimes, i.e., ambient curing and heat curing. The ambient cured

samples in the first group were left in the laboratory environment for 24 hours,

whereas the heat-cured samples in the second group were immediately placed in a

preheated oven and cured at 60 ℃ for 24 hours. All specimens were then demoulded

24 hours after casting. The heat-cured specimens were immediately tested, whereas

the samples in the first group were sealed with cling wrap and kept in a controlled

temperature of 20−23 ℃ and relative humidity of 65 ± 10% until testing at different

ages.

Page 98: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

83

Figure 4.1 Sample preparation: (a) Hobart mixer, (b) mixing process and (c) pouring the

cylindrical moulds

Table 4.2 Mix proportions of one-part geopolymer concrete (kg/m3).

Mix ID Fly ash Slag Activator Water/binder SPf

SSa CHb SOc LHd PCe

M1 1.00 − 0.12 − − − − 0.20 −

M2 1.00 − 0.12 0.120 − − − 0.38 −

M3 1.00 − 0.12 0.09 − − − 0.38 −

M4 1.00 − 0.12 0.09 − 0.006 − 0.38 −

M5 1.00 − 0.12 − 0.090 0.006 − 0.38 −

M6 1.00 − 0.12 0.045 0.045 0.006 − 0.38 −

M7 1.00 − 0.12 0.09 − − 0.06 0.41 −

M8 1.00 − 0.12 0.09 − − 0.10 0.41 −

M9 0.50 0.50 0.12 − − − − 0.30 −

M10 0.50 0.50 0.12 0.09 − − − 0.45 0.009

M11 0.50 0.50 0.12 0.09 − 0.006 − 0.45 0.009

Note: all figures are mass ratios of binder except w/b ratios.

a Sodium silicate (in powder form).

b Calcium hydroxide (in powder form).

c Sodium oxide (in powder form).

d Lithium hydroxide (in powder form).

e Potassium carbonate (in powder form).

f Polycarboxylate ether-based superplasticiser

Page 99: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

84

Testing procedure

The Compressive strength test was conducted at 1 day for the heat-cured samples and

at 3, 7 and 28 days of age for the ambient cured samples using a universal testing

machine. For each age, three cylindrical samples were tested and the reported results

are the average of the three. It should be noted that these heat-cured samples were

not subjected to subsequent ambient curing since previous research works have

demonstrated that the compressive strength of geopolymer is relatively stable after

heat curing (Rangan 2008, Lloyd and Rangan 2010).The loading regime for

conducting the compressive strength tests was according to AS1012.9 (1999). All

cylinders were ground flat on both ends to ensure having smooth and parallel top and

bottom faces prior to performing the tests. The bulk density of the hardened samples

was identified by weighing the 28-day samples.

The microstructure properties of the one-part geopolymer mixes were investigated by

scanning electron microscope (SEM), energy dispersive spectroscopy (EDS),

Fourier-transform infrared spectroscopy (FTIR) and X-ray diffraction (XRD). The

SEM and EDS analyses were conducted using a JEOL 6510LV scanning electron

microscope on freshly-fractured surface of samples with dimensions of about 20

mm× 20 mm ×10 mm which were taken from the identical region of cylindrical

specimens. The FTIR analysis was performed using a Bruker Vertex 70 machine,

which was operated with attenuated total reflection (ATR) method, in 32 scanning

times from 2000 to 400 cm−1 at 4 cm−1 resolution. XRD patterns were obtained by

using Bruker D8 Advance Power Diffractometer. The analysis diffracted within the

range of 10 to 60 2𝜃 using a Cu Kα source (𝜆 = 1.5406 Å). The excitation voltage

was 40 kV at 40 mA with a step size of 0.02º and counting time of 5 s for each step.

Page 100: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

85

The samples of XRD and FTIR analysis were taken from the cylinder specimen and

were ground before the analysis. The XRD samples were mounted to low silicon

background holders and then were placed in instrument.

Results and Discussion

Workability

Mini slump tests were conducted for all the 11 paste mixes and the relative slump

values (Гp) are presented in Figure 4.2 to represent workability. As can be seen, mix

M1 with fly ash and sodium silicate had the highest workability (Гp = 9.3) among

all mixes. By adding calcium hydroxide to mixes M2 and M3, the relative slump

decreased by 42% and 44% compared to that of M1 even though the water to binder

ratio of mixes M2 and M3 was increased by 90%. It is interesting to note that M2

had 33% more calcium hydroxide than M3, but both mixes had almost similar

workability. By introducing lithium hydroxide into M4, its workability increased by

25% relative to that of M3 with a same water/binder ratio. By fully or partially

replacing calcium hydroxide in M4 with sodium oxide, the workability of mixes M5

and M6 increased by 26% and 12%, respectively, compared to that of M4. It should

be noted that noticeable heat release was observed once sodium oxide was mixed

with water because of its exothermic behaviour. Therefore, after mixing, mixes M5

and M6 were left undisturbed for a couple of minutes to allow for cooling down prior

to conducting the mini slump testing. In mixes M7 and M8, additional potassium

carbonate was used as activator. The workability of these two mixes reduced

significantly compared with the reference mix M3 without potassium carbonate.

In mixes M9, M10 and M11, 50 wt% of fly ash was substituted by slag and the

corresponding reference mixes without slag are M1, M3 and M4, respectively.

Page 101: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

86

Adding slag in the mix led to a remarkable decrease in workability which is in

agreement with previous studies (Nath and Sarker 2014). The workability of M9,

M10 and M11 were reduced by 45%, 42% and 49%, respectively, relative to those of

their counterparts. In general, the test results indicate that the use of potassium

carbonate or slag significantly reduces the workability of one-part geopolymers,

whereas sodium oxide and lithium hydroxide have a beneficial effect on the

workability. However, further research is required to identify a clear relationship

between the inclusion of calcium hydroxide and workability of one-part mixes by

considering different variables.

Figure 4.2 Relative slump values of one-part geopolymer mixes

Compressive strength and density

The values of compressive strength at different ages and density at 28 days are

presented in Table 4.3 for different geopolymer mixes. The compressive strength of

all ambient cured samples increases with age. Meanwhile, it is found that the 28-day

Page 102: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

87

compressive strength of an ambient cured mix is very close to that of the heat-cured

counterpart except for mixes M5 and M6. Samples of the two mixes M5 and M6 did

not set under heat curing for 24 hours. This delay in setting might be due to the

accelerated alkali silica reaction that occurred immediately followed by the

exothermic reaction of sodium oxide and water. The reaction of sodium oxide with

siliceous content of silica fume and slag forms silica gel that acts as internal

plasticiser and delays the setting of the mix accordingly. This type of reaction is well

known in glass industry, where sodium oxide or potassium oxide is added to form

silicate to decrease the glass transition temperature (Wallace, Hill et al. 1999).

However, further research is required to address this issue precisely. In general, the

comparison of compressive strengths between heat and ambient cured mixes indicate

that the need of heat curing can be effectively eliminated for these mixes.

As shown in Table 4.3, the measured densities range from 1,680 to 1,940 kg/m3 for

all mixes. It highlights the fact that the type of solid activator has some influence on

the geopolymer density. Compared with the lowest density of 1,680 kg/m3, there is

an increase of 15.5% for the highest density of 1,940 kg/m3. However, there is no

clear correlation between the density and 28-day compressive strength (fc). For

example, mix M1 has the highest density of 1,940 kg/m3 but the fc-value at 28 days

is only 10.9 MPa. In contrast, mix M11 has the lowest density of 1,680 kg/m3 and the

highest strength of 38.0 MPa.

Page 103: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

88

Table 4.3 Test results of geopolymer concrete mixes.

Mix ID Compressive strength (MPa) Density

(kg/m3) Heat cured Ambient cured

3 d 7 d 28 d

M1 12.3±0.1 N/Aa 7.0±1.3 10.9±1.1 1,940

M2 18.3±0.8 8.2±0.5 13.7±0.3 18.1±1.4 1,700

M3 16.0±0.2 6.6±0.4 12.3±0.1 15.2±0.5 1,680

M4 23.0±1.1 12.2±0.1 17.3±0.8 22.6±0.7 1,730

M5 N/Aa N/Aa 6.3±0.4 31.2±0.9 1,940

M6 N/Aa N/Aa 4.0±1.2 17.5±1.0 1,790

M7 10.0±0.4 4.2±1.2 5.2±2.0 9.9±1.1 1,680

M8 11.2±2.1 4.9±0.4 6.5±0.3 11.6±0.7 1,740

M9 23.0±0.6 9.7±1.3 23.8±0.7 24.3±1.2 1,760

M10 27.5±0.8 16.6±0.2 22.0±0.5 27.2±1.0 1,730

M11 39.3±0.5 20.3±0.8 28.6±0.4 38.0±0.8 1,680

a The mixture did not set and no strength data was recorded.

Mix M1 using only sodium silicate as activator has a relatively low strength. This

mix did not set within 24 h after casting and did not have any strength even at 3 days.

A 28-day compressive strength of 10.9 MPa was obtained for this mix. By adding

Ca(OH)2 to mixes M2 and M3, an obvious strength increase, especially at early ages,

is obtained compared with the reference mix M1. This is expected as the increase in

alkalinity could promote the dissolution of aluminate and silicate ions from the fly

ash (Temuujin, Van Riessen et al. 2009). Since M2 has more Ca(OH)2 (12 wt%) in

the mix than M3 (9 wt%), fc of M2 at 28 days (18.1 MPa) is 19% higher than that of

M3 (15.2 MPa).

By introducing additional lithium hydroxide into mix M4, fc of this mix at 28 days

increases further to 22.6 MPa compared to 15.2 MPa of the reference mix M3. By

fully replacing calcium hydroxide in M4 with sodium oxide in mix M5, a 28-day

compressive strength of 31.2 MPa was achieved, representing a strength increase of

Page 104: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

89

38% relative to that of M4. But when only half of calcium hydroxide in M4 was

replaced by sodium oxide in mix M6, no obvious benefits are observed. In fact, the

strength of M6 decreases at all ages compared with the corresponding strength of the

reference mix M4. Further research can be conducted to determine the optimised

amount of sodium oxide in the mix design and propose methods to accelerate setting

of geopolymer with sodium oxide. Furthermore, the extent of heat from the

exothermic dissolution/hydration reaction of sodium oxide with water should be

investigated in more detail and if there could be health issues and if upscaling could

cause certain problems in terms of heat development as high content NaOH solutions

tend to boil during preparation.

Based on the mix design of M3, additional potassium carbonate was introduced into

mixes M7 and M8. However, this yields no benefits in the modified mixes. On the

contrary, there is a strength decrease for the two mixes at all ages compared with the

reference mix M3. The decelerating function of alkaline carbonates in the presence

of other activators was also observed previously (Fernández-Jiménez and Palomo

2005, Fernández-Jiménez, Palomo et al. 2006). The inclusion of carbonate reduces

the release of Al and Si from fly ash and promotes the condensation of the species

and polymerisation of the aluminosilicate gel. However, as the rate of the reaction

decreases quickly due to the high decrease of the system pH, low mechanical

strength and a very porous structure is achievable. Slag along with fly ash was used

as geopolymer precursor in mixes M9, M10 and M11. Compared with the reference

mixes M1, M3 and M4 without slag, the inclusion of slag in geopolymer results in

68−123% increase in 28-day compressive strength. Meanwhile, the compressive

strengths at 3 and 7 days are also increased significantly. This increase in

compressive strength by inclusion of slag in binder was also observed in

Page 105: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

90

conventional geopolymer mixes. For example, Nath and Sarker (Nath and Sarker

2014) reported an increase of almost 10 MPa in the 28-day compressive strength for

every 10% increment of the slag content.

As shown in Table 4.3, five out of 11 mixes have 28-day compressive strengths over

20 MPa and Mix 11 has achieved the highest strength of 38.0 MPa. Despite the fact

that the type/combination of solid activators has significant influence on the

properties of geopolymers, the increase of dosage does not necessarily have positive

effect. Mixes M7 and M8 with considerable amount of solid activators (27% and

31% of binder content, respectively) not only did not achieve reasonable mechanical

strength, but also increased the cost and possible hazards which in turn covers the

main advantage of developing one-part geopolymers. One of the issues associated

with excess of alkalis in some conventional geopolymer formulations is

efflorescence especially in case of ambient cured geopolymers. On the other hand,

mixes M4 and M11 which consisted 21% of binder, solid activators, gained

promising strength at 28 days of age. Therefore, still more investigation is required to

propose more cost effective and environmental friendly one-part geopolymers by

minimising concentration of activators.

Although the achieved strengths in the current study are not very high, it is still

higher than the results of many conventional geopolymer paste mixes (Somna,

Jaturapitakkul et al. 2011, Saha and Rajasekaran 2017, Pan, Tao et al. 2018). It is

possible to use the developed binders to make geopolymer concrete for some

particular applications, such as non-structural concrete or structural concrete with a

low strength requirement, even though by adding aggregate to these geopolymer

pastes, higher mechanical strength is achieved even at ambient temperature curing.

Page 106: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

91

Therefore, further research should be focused on the evaluation of the structural

performance of one-part geopolymer concrete elements.

FTIR Results

The infrared spectra of raw materials and typical ambient cured fly ash-based

geopolymer mixes (M1, M4 and M5) and fly ash/slag-based mixes (M9, M10 and

M11) are illustrated in Figure 4.3. The geopolymer samples were cured for 28 days

before conducting the tests. As can be seen, the FTIR spectra of fly ash-based mixes

are different from those of the fly ash/slag-based mixes in terms of band shape and

position. Previous studies (Puertas, Martınez-Ramırez et al. 2000, Garcia-Lodeiro,

Palomo et al. 2011, Yang, Yao et al. 2012, Abdalqader, Jin et al. 2016,

Hajimohammadi and van Deventer 2017, Kaja, Lazaro et al. 2018) highlight that the

major vibrational band in the range of 950-1100 cm−1 associated with anti-symmetric

Si-O-T (T=Si or Al) stretching vibrations is very useful to understand the network

structure of geopolymer, especially the polymerisation degree. The shifting of the T-

O band shows that the silicates and alumina-silicates in the precursor have

participated in the reaction and formed aluminosilicate geopolymer gels to develop

mechanical strength (Samantasinghar and Singh 2018). For fly ash-based mixes, this

band position changed from 1040 cm−1 in the raw fly ash to 1025, 1000 and 1000

cm−1 in M1, M4 and M5, respectively. The lower wavenumbers of mixes M4 and M5

compared to M1 indicate a higher degree of silicon substitution by aluminum, which

is due to the increased alkalinity in M4 and M5. This explains the higher

compressive strength of M4 and M5 compared to M1. The fly ash/slag-based mixes

have lower wavenumbers of T-O band compared to fly ash-based mixes. This is

consistent with previous research findings (Pan, Tao et al. 2018) regarding the

Page 107: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

92

influence of adding slag in fly ash-based geopolymer. The decrease in wavenumbers

by the inclusion of calcium dissolved from the slag could be due to either an increase

in concentration of tetrahedral Al substitution or a decrease in network connectivity

resulting from lower polymerisation degree (Yang, Yao et al. 2012). The T-O band

was originally at around 920 cm−1 in the raw slag and was shifted towards higher

wavenumbers in the fly ash/slag-based mixes due to higher polymerised Si-O

network (Gao, Yu et al. 2015). The wavenumber assigned to the maximum peaks of

M9, M10 and M11 are 958, 952 and 952 cm−1, respectively. These results reveal that

the mix structure is less polymerised after adding slag and calcium hydroxide. In

previous studies, the absorption bands in the range of 950-980 cm−1 were attributed

to the formation of C-A-S-H gel with short chain, whilst the bands in the range of

1000-1100 cm-1 are assigned to the formation of N-A-S-H gel (Rees, Provis et al.

2007, Garcia-Lodeiro, Palomo et al. 2011, Ismail, Bernal et al. 2013, Gao, Yu et al.

2015). The formation of C-A-S-H type gel decreases the Si and Al concentration and

consumes the available alkali cations for geopolymerisation process. Due to the high

calcium content in the fly ash/slag-based mixes, usually insufficient Si and Al remain

to promote the formation of high polymerised structures after consumption of

calcium (Gao, Yu et al. 2015). Thus, it can be postulated that the main reaction

product in the fly ash-based mixes is N-A-S-H gel, whereas that in the fly ash/slag-

based mixes is C-A-S-H gel.

In the FTIR spectra of fly ash/slag-based mixes, another absorption band at 1068 cm-

1 is observed, which results from the partial substitution of Si and Al in the gel

structure (Puertas, Martınez-Ramırez et al. 2000). The bands at 880 and 1400 cm−1

indicate the formation of carbonates, which could result from the presence of calcite

as also identified by XRD (Abdalqader, Jin et al. 2016). The weak band at 1640-1650

Page 108: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

93

cm−1 represents the stretching of O-H groups and bending vibration of water

molecules (Peng, Wang et al. 2015). All the spectra in Fig. 3 exhibit a peak in the

range of 420-440 cm−1, which is associated with SiO4 tetrahedral deformation

(Sakulich, Anderson et al. 2009). It should be mentioned that FTIR spectra for heat

cured mixes are almost identical to those of the ambient cured ones shown in Figure

4.3, which is consistent with previous research findings (Puertas, Martınez-Ramırez

et al. 2000, Puertas and Fernández-Jiménez 2003).

Figure 4.3 FTIR spectra of raw materials and typical geopolymer mixes: (a) fly ash-based

geopolymers; and (b) fly ash/slag based geopolymers.

XRD patterns

The XRD patterns of raw fly ash and raw slag, fly ash-based mixes, fly ash/slag-

based geopolymer mixes and comparison of ambient temperature curing versus heat

curing for M10 are exhibited in Figure 4.4. It should be mentioned that according to

the results of this experiment and consistent with previous studies (Fernández-

Jiménez and Palomo 2005, Bernal, Provis et al. 2015, Abdalqader, Jin et al. 2016,

Pan, Tao et al. 2018), the major and important phases for geopolymers were

observed within the range of 10 to 60 2𝜃, and accordingly this range was selected.

Page 109: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

94

Both fly ash and slag consist of mainly semi-crystalline and amorphous phases. The

main crystalline phases in fly ash are quartz, mullite and a small amount of

maghemite, which remain unreacted in fly ash-based geopolymer. The raw slag has

crystalline phases associated with aragonite, gypsum, calcite and vaterite. The

gypsum was introduced in the milling process of slag.

In general, all geopolymer mixes have a broad diffuse hump in the range of 20-40º

2𝜃 which highlights the amorphous feature of the geopolymeric products. The XRD

pattern of fly ash-based mix is very similar to that of raw fly ash. No new crystalline

phase has formed due to alkali activation at room temperature which is in agreement

with previous studies (Somna, Jaturapitakkul et al. 2011, Pan, Tao et al. 2017). The

XRD pattern of the fly ash/slag-based mix is different from that of the fly ash-based

samples. For the fly ash/slag-based mix, a peak at around 29.5º 2𝜃 is identified,

showing the presence of C-S-H gel in the reaction product together with traces of

calcite. As also reported by others (Xu, Provis et al. 2008, Bernal, Provis et al. 2015,

Abdalqader, Jin et al. 2016, Samantasinghar and Singh 2018), the presence of calcite

is from the recarbonation of Ca from absorbing CO2 from CO32− ions and other

calcium carbonate polymorphs in slag such as vaterite and aragonite. However, the

nature of C-S-H formed in these mixes is different from the normal CSH gel formed

in OPC due to the lower CaO/SiO2 molar ratios of these mixes (on average 0.75 for

fly ash/slag mixes compared to a mean of 1.75 for neat OPC paste) (Yip and Van

Deventer 2003). Depending on the calcium content, previous studies (Kumar, Kumar

et al. 2010, Nath and Kumar 2013, Pan, Tao et al. 2017) reported the formation of C-

A-S-H gel or a coexistence of N-A-S-H and C-A-S-H gels after the inclusion of

calcium compounds in fly ash-based geopolymers. The formation of C-A-S-H gel in

fly ash based geopolymers with 50wt% slag inclusion, as also reported previously

Page 110: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

95

(Nath and Sarker 2014, Pan, Tao et al. 2017) helps the development of early and

ultimate strength.

Another finding from the current study is that the XRD pattern of geopolymer is not

obviously affected by the activator content or curing condition. However, the

activator content affects the intensity of the formed phases, and in turn the final

hydration products, although the amorphous nature of the geopolymer binders

making it difficult to recognise the actual difference between these products. In the

fly ash-based mixes, the inclusion of lithium hydroxide together with calcium

hydroxide and/or sodium oxide in mixes M4−M6 caused a reduction in the quartz

peak and hence an increase in the amount of hydration products. This agrees well

with the test results of strength. Also, in the fly ash/slag-based mixes, M11 presents

the highest intensity in CSH peak, which justifies the achievement of the highest

compressive strength among these mixes.

Figure 4.4 XRD patterns of : (a) raw fly ash and raw slag, (b) ambient cured fly ash-based

geopolymers, (c) ambient cured fly ash/slag based geopolymers and (d) ambient cured

against heat cured M10. (Aragonite-A, Calcite-C, Gypsum-G, CSH-□, Quartz-Q, Mullite-M,

Maghemite-Mg).

Page 111: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

96

SEM/EDS results

Figures 4.5 to 4.7 present the characteristic morphology of selected geopolymer

mixes at 28 days of age using backscattered electron (BSE). It should be noted that

lithium is too light to be identified by SEM/EDS analysis, so in the mixes containing

lithium hydroxide, the lithium content was not presented. In general, the

microstructure of different mixes showed different morphologies associated with the

inclusion of slag and type/combination of solid activators.

Figure 4.5 depicts the BSE images of mixes M1, M4 and M5 in which fly ash was

used as the sole geopolymer precursor. The EDS results show the products have high

content of Si and Al but negligible amount of Ca in mixes M1 and M5 and a small

amount of Ca in M4 due to the inclusion of calcium hydroxide in the activator. The

EDS results confirm that the main reaction product in fly ash-based mixes is N-(C)-

A-S-H geopolymer gel in the binding phase. Mix M1 using only sodium silicate as

activator has a porous structure with a huge amount of unreacted fly ash in the

matrix. The incomplete geopolymerisation leads to low compressive strength. Mix

M5 presents more compact and dense morphology than M1 and M4. The former has

less unreacted fly ash in the matrix, indicating the addition of lithium hydroxide and

sodium oxide in M5 with sodium silicate results in better microstructure and

morphology. It should be mentioned that M5 has many microcracks in its

microstructure, which might be due to temperature cracks caused by the heat

generated from the reaction of sodium oxide with water.

Figure 4.6 depicts the micrographs of mixes M9, M10 and M11 containing 50 wt%

fly ash and 50 wt% slag in the binder. These micrographs show that the samples have

high content of Ca as identified by the EDS analysis. It has been widely reported that

Page 112: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

97

the matrix phase formed in fly ash/slag-based geopolymer is not a three-dimensional

structure, but is rather similar to C-S-H gel with linear chains of Si-O-Si (Yang, Yao

et al. 2012). This justifies the variations of pattern in XRD and wavenumber of T-O-

Si bands in FTIR spectra compared with conventional fly ash-based geopolymer. The

calcium in slag can reduce polymerisation degree, which is beneficial for the pore

size distribution and strength development (Duxson, Fernández-Jiménez et al. 2007,

Yang, Yao et al. 2012). It is worth noting that the formed C-S-H gel in fly ash/slag-

based geopolymer is rich in Al and has a low Ca/Si ratio of ~ 0.6 (Yip, Lukey et al.

2005). In contrast, the hydration product of OPC has a much higher Ca/Si ratio of

1.5-1.7.

In fly ash/slag-based geopolymer mixes, unreacted fly ash and slag particles are also

found. However, the fly ash/slag-based geopolymer mixes have denser

microstructure than fly ash-based mixes. This is further illustrated in Figure 4.7 by

comparing the morphologies of mixes M1 (fly ash) and M11 (fly ash/slag) taken in

secondary electron mode. It is not surprising that M1 has a very low compressive

strength due to the very porous matrix and a huge amount of unreacted fly ash

particles. On the other hand, M11 presents very dense and compact matrix and as a

result has high compressive strength.

Page 113: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

98

Figure 4.5 BSE images of fly ash based one-part geopolymer mixes.

Figure 4.6 BSE images of fly ash/slag based one-part geopolymer mixes.

Page 114: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

99

Figure 4.7 SEM images of geopolymer mixes of M1 and M11.

Conclusions

In this chapter, one-part geopolymer mixes have been successfully developed using

100% fly ash or a mixture of 50 wt% fly ash and 50 wt% slag as the precursor.

Different solid activators, such as sodium silicate, calcium hydroxide, sodium oxide,

lithium hydroxide and potassium carbonate or their combinations, have been used to

replace highly corrosive alkali solutions in conventional geopolymer. The following

conclusions can be drawn within the scope of this chapter.

1. The type of aluminosilicate precursor has substantial influence on the

geopolymerisation development and accordingly the final geopolymer gel formed.

As confirmed by the SEM/EDS, XRD and IR analyses, the reaction product in fly

Page 115: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

100

ash-based mixes is mainly dominant by N(C)-A-S-H type gel. However, in the fly

ash/slag-based mixes, the reaction product is C-S-H type gel rich in Al. The inclusion

of slag in the mixture results in more compact and denser structure with less

unreacted particles compared to the fly ash-based mixes. Accordingly, the fly

ash/slag-based mixes develop higher early and final strength but have lower

workability.

2. The type/combination of solid activators also affect the mechanical and

microstructure properties of one-part geopolymer mixes. While sodium silicate is

used as the sole solid activator, a weak and porous structure is achieved, and the

compressive strength of the geopolymer is low. The inclusion of calcium hydroxide

in activator results in a substantial increase of compressive strength but leads to a

decrease in workability. The combined use of sodium silicate, calcium hydroxide and

potassium carbonate produced a mix with very low strength and workability. A

relatively high compressive strength is obtained when sodium silicate, calcium

hydroxide (or sodium oxide), and lithium hydroxide are combined and used as

activator. However, the use of sodium oxide leads to very slow setting of

geopolymer.

3. When a small amount of lithium hydroxide along with calcium hydroxide and

sodium silicate is used as activator, a 28-day compressive strength of 22.6 MPa is

achieved for fly ash-based geopolymer and 38 MPa is achieved for fly ash/slag-based

geopolymer. The developed binders can be used to make ambient cured geopolymer

concrete for some particular applications, such as non-structural concrete or

structural concrete with a low strength requirement.

Page 116: Mechanical, Thermal and Structural Investigations of One

Chapter 4 Development of One-Part Geopolymers

101

Further research should be conducted to determine the optimised dosage of lithium

hydroxide in combination with other activators, such as calcium oxide, sodium

aluminate, calcium aluminate and lithium carbonate, to further improve the

compressive strength. Meanwhile, mechanical properties, long-term performance,

durability and efflorescence of one-part geopolymer concrete using the developed

binders should also be investigated.

Page 117: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

102

Chapter 5 Mechanical and Thermal Properties

General

Mechanical and thermal properties of ambient cured one-part hybrid OPC-

geopolymer concrete (HGC), one-part geopolymer concrete (OGC) and conventional

OPC concrete (CC) have been evaluated in the laboratory as part of this research. It

should be noted that HGC and OGC mixes in this chapter are chosen based on the

optimised mixes derived in Chapters 3 and 4.

Mechanical properties including compressive strength, modulus of elasticity (MOE),

tensile strength (Brazil split), modulus of rapture (MOR) and stress-strain

relationship curve were measured at ambient temperature for all types of materials.

Furthermore, compressive strength and strain at peak stress at elevated temperatures

were measured for concrete mixes.

Thermal properties including thermal conductivity and specific heat were assessed

for all types of materials.

This Chapter includes the introduction and description of materials, testing

procedures, results and discussion of the tests followed by the conclusion of the

chapter.

Experimental Procedure

Materials

The low calcium fly ash (ASTM C 618 Class F) used in this research was sourced

from Gladstone Power Station, Queensland, Australia. Commercially available slag

Page 118: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

103

supplied by Australian Builders and OPC (ASTM C 150 Type 1) supplied by Cement

Australia were also used in the experiments. The results of X-ray Fluorescence

(XRF) analyses of the raw materials are presented in Table 5.1. Natural river sand

with a specific gravity of 2.6 and water absorption of 3.5% was used as fine

aggregate, whereas crushed basalt with a maximum nominal size of 10 mm, a

specific gravity of 2.8 and water absorption of 1.6% was used as coarse aggregate.

Prior to mixing, all aggregates were dried in an oven at 105℃ for a period of 48

hours to remove any moisture.

Table 5.1 Chemical composition of OPC, fly ash and slag determined by XRF

Material SiO2 Al2O3 Fe2O3 CaO MgO Na2O K2O SO3 TiO2 LoIa

OPC (%) 20.2 4.9 2.6 62.7 2.0 0.2 0.4 2.2 − 2.0

Fly ash (%) 50.3 27.56 12.68 2.7 1.3 0.6 0.7 0.36 1.42 0.67

Slag (%) 32.5 13.0 0.55 43.2 4.7 0.2 0.3 4.3 0.51 0.08

a Loss of ignition

For HGC mix, powder form potassium carbonate denoted as “K2CO3” with 99.99%

purity was used as activator and was supplied by Sigma-Aldrich in Australia. For

OGC mix, a combination of three activators namely: sodium silicate, calcium

hydroxide and lithium hydroxide was used. The sodium silicate-based activator

denoted as “Na2SiO3-GD” was supplied by PQ Australia. The supplied sodium

silicate has the chemical composition: 27% Na2O, 54% SiO2 and 19% H2O with a

modulus ratio (Ms = SiO2/Na2O) of 2.0. An industrial grade calcium hydroxide

Ca(OH)2 commonly used in the construction industry was supplied by Orica

Chemicals in Australia. A Reagent grade lithium hydroxide denoted as “LiOH” with

98% purity was supplied by Sigma-Aldrich, Australia.

Page 119: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

104

Mix proportions and mixing

A total of 3 mixes including one-part hybrid OPC-geopolymer concrete (HGC), one-

part geopolymer concrete (OGC) and their control conventional OPC concrete (CC)

were investigated in this study. Details of the mix proportions are presented in Table

5.2. As previously mentioned in Chapter 2, thermal properties of concrete is

significantly influenced by its moisture content. Therefore, the water to binder ratio

for all mixes was kept constant at 0.45. Also, all of the mixes had the same amount

of binder and aggregate. The aggregate proportions shown in Table 5.2 were based

on the saturated surface dry (SSD) condition.

In one-part geopolymer concrete mix (OGC), binder consisted of 50 wt.% of fly ash

and 50 wt. % of slag activated by 12 wt.% of the binder sodium silicate, 9 wt.% of

the binder Ca(OH)2 and 6 wt.% of the binder lithium hydroxide. The binder of the

one-part hybrid OPC-geopolymer concrete mix (HGC) comprised 30 wt.% of the

total binder OPC, 63 wt.% fly ash and 7 wt.% slag and was activated using solid

potassium carbonate with a weight of 7.5% of the total geopolymeric raw materials

(fly ash and slag).

Table 5.2 Details of mix proportions (kg/m3)

Mix

No. Designation Coarse

aggregate Sand OPC Fly ash Slag Activator Citric acid Water/binder Superplasticiser

1 CC 1220 635 400 - - - - 0.45 0.5

2 HGCa 1220 635 120 252 28 21 0.48 0.45 1.0

3 OGCb 1220 635 - 200 200 86.4 - 0.45 3.0

a Potassium carbonate (powder form) was used as activator

b Sodium silicate, calcium hydroxide and lithium hydroxide (powder form) were combinedly

used as activator

Page 120: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

105

All mixes were prepared using a vertical pan mixer. Solid ingredients including sand

and coarse aggregate together with OPC in CC, OPC, fly ash, slag, activator and

citric acid in HGC and fly ash, slag and activators in OGC were added to the mixer

and dry-mixed thoroughly for 3 min before adding water. The water and

superplasticiser were then added gradually and mixing was continued for another 4-6

min until a uniform mix was achieved (Figure 5.1). Then, the slump was adjusted to

100±10 mm by addition of superplasticiser. The freshly mixed concrete was then

cast in 100 × 200 mm cylindrical and 100 × 400 mm prism moulds in two layers and

each layer was compacted on a vibrating table to achieve proper consolidation. A

total of 80 cylinders and 3 prisms for each types of concrete were cast. The samples

were demoulded 24 hours after casting (Figure 5.2) and were wrapped with plastic

sheet. Then they were stored in a controlled environment with temperature of

20−23 ℃ and relative humidity of 65±10% until testing.

Figure 5.1 Mixing concrete in laboratory

Page 121: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

106

Figure 5.2 Demoulded specimens after 24 h

Testing procedure

5.2.3.1 Mechanical properties at ambient temperature

Compressive strength, modulus of elasticity (MOE), indirect tensile strength (Brazil

split), modulus of rapture (MOR) and stress-strain relationship curve were measured

at 28 days of age for OGC, HGC and CC at ambient temperature in accordance with

the relevant Australian Standards (AS1012 series). For each test, three cylindrical or

prismatic samples were tested and the reported results are the average of the three.

Prior to performing the tests, all the cylinders were properly capped to ensure having

smooth and parallel top and bottom faces.

All compressive strength tests were performed on cylindrical specimens at a loading

rate of 20 ± 2 MPa/min according to AS 1012.9 (1999) using a universal testing

machine. The compressive strength of the specimens was measured by dividing the

Page 122: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

107

maximum applied force to the specimen over the cross sectional area. Modulus of

elasticity tests (MOE) were performed according to AS 1012.17 (1997). Static chord

(MOE) is defined as gradient of the chord drawn between two specific points on the

stress-strain curve as follows:

Point 1: is 50 micro-strains and stress corresponding to this strain and Point 2: is the

stress equivalent to “test load” and its corresponding strain, where the test load is

equal to 40 percent of the average compressive strength, tested in accordance with

AS 1012.9 immediately prior to the static chord MOE test. This test is conducted

under a load control condition with a loading rate equivalent to 15 ± 2 MPa per

minute. The MOE of the concrete samples can be determined from following

Equation 5.1:

𝐸𝑐 =(G2 − G1)

(ε2 − 0.00005) (5.1)

where Ec is the concrete modulus of elasticity in MPa, G2 is the test load mentioned

above over cross-sectional area of the specimen, G1 is the applied load at a strain of

50 × 10-6 over the cross-sectional area of the specimen and ε2 is strain associated

with deformation at test load equal to deformation over the length of the gauge in

10-6 m/m. The test setup for MOE test is shown in Figure 5.3.

Page 123: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

108

Figure 5.3 MOE test setup

The indirect tensile strength (Brazil or splitting test) of concrete mixes were

performed according to AS 1012.10 (2000). This test is a simple and indirect way of

measuring the tensile strength of concrete and the strength achieved is believed to be

equivalent to the direct tensile strength of concrete (Carmona and Aguado 2012). The

Brazilian test involves subjecting a cylindrical specimen to diameter compression by

applying the load uniformly along a line, on two diametrically opposite generatrices

as shown in Figure 5.4 until failure. The maximum load is recorded and the indirect

tensile strength of the specimen is measured using Equation 5.2;

𝑓𝑠𝑡 =2000𝑃

𝜋𝐿𝐷 (5.2)

where, 𝑓𝑠𝑡 is indirect tensile strength in MPa, P is the maximum applied load in kN,

L is the length of specimen in mm and D is the dimeter in mm.

Page 124: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

109

(a) (b)

Figure 5.4 Splitting tensile test; (a) typical test setup and (b) specimen after failure

Modulus of rapture (MOR) test was conducted on prism specimens for all concrete

mixes by applying 4-point flexural load at a loading rate of 1 ± 0.1 MPa/min until

failure according to AS 1012.11 (2000). The test setup for MOR test is depicted in

Figure 5.5. Flexural strength is expressed in terms of the MOR, which is the

maximum stress at rapture determined from the following Equation 5.3.

𝑓𝑐𝑡.𝑓 =𝑃𝐿1000

𝐵𝐷2 (5.3)

where 𝑓𝑐𝑡.𝑓 is the MOR in MPa, P is the maximum applied force in kN, L is span

length in mm, B is the average width of the specimen at the section of failure in mm

and D is the average depth of specimen at the failure section in mm.

Page 125: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

110

Figure 5.5 Test setup for MOR test

The compressive stress-strain behaviour of concrete specimens was also measured.

The cylindrical specimens were subjected to uniaxial compressive load and the axial

and circumferential strain were recorded using extensometer/strain gauges attached

to the specimen as shown in Figure 5.6. Testing was performed under closed-loop

displacement control which allows for more controlled evaluation of the post-failure

strain response. Specimens were loaded in pure uniaxial compression at a constant

displacement rate of 0.2 mm/min and were instrumented with an axial and radial

extensometers.

Page 126: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

111

Figure 5.6 Test setup for compressive stress-strain test

5.2.3.2 Mechanical properties at elevated temperatures

The compressive strength and strain at peak stress of OGC, HGC and CC were

measured at elevated temperatures of 200 ℃, 400 ℃, 600 ℃ and 800 ℃ according to

RILEM TC 129-MHT (1995) procedures. For this purpose, cylindrical specimens

without pre-loading were subjected to a very slow heating rate of 4 ℃/min in order to

reproduce quasi-steady thermal conditions. The tests were carried out in an electric

split-tube furnace equipped with a MTS actuator as shown in Figure 5.7. After

reaching the target temperature, the samples remained at the target temperature for

Page 127: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

112

further 1.5 hours to assure the quasi-steady condition has been reached. Then the

samples were loaded up to failure. For each target temperature, three cylindrical

specimens were tested and the reported results are the average of the three.

Figure 5.7 Test setup for investigation of compressive strength of samples at elevated

temperatures

5.2.3.3 Thermal properties

The thermal properties of concrete mixes at ambient and elevated temperatures

including thermal conductivity and specific heat were simultaneously determined in

this study using transient plane source (TPS) method. According to this procedure, a

flat round hot disc sensor is placed between two pieces of material. Then the sensor

consists of a thin Nickel foil spiral which is compressed between two sheets of

electrical insulation material (Jansson 2004). In this study, the insulation material

used for ambient temperature testing was Kapton with a thickness of 25 𝜇𝑚 and

Mica with a thickness of 60 𝜇𝑚 for elevated temperature testing. The test setup for

Page 128: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

113

ambient and elevated temperature testing according to TPS procedure is shown in

Figure 5.8. The hot disc acts as a constant effect generator and a resistance

thermometer at the same time. The time dependant resistance rise is recorded and

converted with the temperature coefficient of resistivity for Nickel to a temperature

response curve. While a constant electrical effect is applied, the temperature in the

sensor rises and heat starts to flow to the tested material. The temperature rise in the

sensor is then a direct response of the thermal properties of the tested material. The

size of sensor, measurement times and suitable applied effect should be selected

based on the thermal properties, so it is an iterative process if the properties of the

tested material are totally unknown (Log and Gustafsson 1995). The disc samples (of

65 mm in diameter and 25 mm in length) were sliced from the mid-section of

cylindrical samples before the measurement. To define thermal properties at elevated

temperatures, the apparatus was connected to a tube furnace in which a specimen

was heated to a predetermined target temperature. These series of tests were started

at the age of 120 days. In this study the test procedure proposed by Pan et al. (Pan,

Tao et al. 2018) was used. Firstly, the specimens were dried in an oven at 105 ºC for

72 hours before being immediately subjected to heating in the tube furnace. In each

test, the furnace temperature was increased to the lowest target exposure temperature

(100 ºC) and maintained at that level for at least 10 hours. then the specific heat and

thermal conductivity were measured followed by increase of temperature up to 600

ºC in the furnace and this process was repeated.

Page 129: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

114

Figure 5.8 The thermal properties measurement according to TPS; (a) test specimens, (b)

general view of the test setup, (c) test setup for ambient temperature testing and (d) test setup

for elevated temperature testing

Test Results and Discussion

In this section, the results of mechanical and thermal properties tests mentioned

earlier in the chapter will be discussed. Comparison is also made between the

characterisation of one-part geopolymer concrete (OGC), one-part hybrid OPC-

geopolymer concrete (HGC) and OPC concrete.

Mechanical properties at ambient temperature

The mechanical properties of OGC, HGC and CC concrete including compressive

strength, MOE, MOR and splitting tensile strength at 28 days are shown in Table 5.3.

The compressive strength of OGC and HGC which had the same aggregate and

Page 130: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

115

binder content and also identical water to binder ratio compared to CC, were in the

range of 40 MPa compared to 50 MPa for CC mix. The mean value of MOE for

OGC, HGC and CC were 24.5, 28.1 and 32.8 GPa, respectively. The results of

flexural strength (MOR) were higher than values for the indirect tensile strength

which is consistent with previous findings (Nath and Sarker 2017). The OGC and

HGC mixes with the same compressive strength grade showed almost similar tensile

behaviour but both had lower values compared to CC.

Table 5.3 Mechanical properties of concrete mixes

Designation Compressive

strength (MPa)

MOE

(GPa)

MOR

(MPa)

Splitting

tensile (MPa)

CC 52.5±0.3 32.8 4.8 4.1

HGC 42.2±1.0 28.1 4.0 3.6

OGC 40.4±0.7 24.5 3.6 3.4

The stress-strain curves of OGC, HGC and CC are illustrated in Figure 5.9. The

mixes presented different behaviour in the ascending branch up to the peak stress and

also after failure. The stress-strain curve for mix CC was steeper and more linear on

the ascending branch than that of HGC and OGC. The strain at peak stress for OGC

and HGC was almost similar but higher than that of CC. Consistent with previous

studies, increase of OPC in CC compared to OGC and HGC, enhanced the

compressive strength and increased the MOE but decreased the strain capacity at

peak stress (Phoo-ngernkham, Chindaprasirt et al. 2013). In terms of post-peak

behaviour, OGC and HGC demonstrated different patterns than that of CC. Where

CC showed gradual strain softening beyond the ultimate stress, OGC and HGC

displayed brittle fracture with little or no softening following the peak. This high

Page 131: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

116

quasi-brittle behaviour of geopolymers which was also observed previously (Thomas

and Peethamparan 2015, Noushini, Aslani et al. 2016), has been generally attributed

to their ceramic-like nature.

Figure 5.9 Stress-strain behaviour of concrete mixes after 28 days

One convenient way to quantify the shape of different concrete is to normalise the

stress-strain curve. Average stress-strain curves were normalised by their peak stress

and the strain at peak stress. The normalised stress-strain curves for OGC, HGC and

CC are compared with each other in Figure 5.10. As can be seen, OGC and HGC

mixes displayed almost similar behaviour at ascending branch but different from that

of CC mix. The descending branch of concrete normally indicates the ductility of

concrete. The variation of normalised stress-strain curve at descending branch

between different types of concrete became significant. In general, it was found that

OGC mix showed the least ductility among others while CC and HGC had gentler

descending branch than OGC. These stress-strain curves will be compared with the

0

10

20

30

40

50

60

0 2000 4000 6000 8000

Stre

ss (

MPa

)

Strain (με)

OGC HGC CC

Page 132: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

117

current concrete models for stress-strain relationship curves at the end of this chapter

to evaluate if the current models can be used for these types of concrete or not.

Figure 5.10 Comparison of normalised stress-strain curves for different concrete mixes

Mechanical properties at elevated temperatures

The hot strength and strain at peak stress of OGC, HGC and CC samples were

measured at the age of 120 to 150 days after heating the specimens in quasi-steady

condition. This approach was applied to ensure the risk of dangerous thermal self-

stresses is minimized (Pan, Tao et al. 2017). The reduction factor of hot compressive

strength of OGC, HGC and CC samples at temperatures of 23, 200, 400, 600 and 800

°C are shown in Figure 5.11. All concrete mixes of OGC, HGC and CC showed

similar pattern from 23 to 200 °C and experienced compressive strength deterioration

while beyond this range they showed distinct behaviour. The decrease in

compressive strength of mixes from ambient temperature is reported to be due to a

number of temperature-induced stresses and chemical reactions (Shaikh and

0

0.2

0.4

0.6

0.8

1

1.2

0 1 2 3 4

No

rmal

ised

str

ess

Normalised strain

OGC HGC CC

Page 133: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

118

Vimonsatit 2015). This strength loss mainly results from different thermal

incompatibilities between the paste and aggregate of the concrete while by increasing

temperature aggregate expands and paste contracts due to loss of moisture up to

about 350 °C (Kong and Sanjayan 2010). This thermal incompatibility causes

cracking and damaging the bond between the aggregate and paste in concrete. The

other reason associated with compressive strength loss is the thermal gradient

between the surface and core temperatures of cylinders which is greater in lower

heating temperatures (Shaikh and Vimonsatit 2015). For OGC specimens, the

increase of strength about 32% from 200 to 600 °C was observed while the strength

was still lower than ambient one. The compressive strength in HGC samples increase

by 36% from 200 to 400 °C followed by 20% drop from 400 to 600 °C. The strength

increase in OGC and HGC specimens might be attributed to the stable contraction of

geopolymer paste (Provis, Lukey et al. 2005, Shaikh and Vimonsatit 2015). At 800

°C, significant loss of strength was observed for OGC and HGC samples which is

consistent with previous findings and results from increase in geopolymer paste

shrinkage due to viscous sintering of geopolymer matrix filling the voids in the

material (Rickard, Riessen et al. 2010, Rickard, Temuujin et al. 2012). On the other

hand, the loss in compressive strength continued for CC samples after 200 °C. The

substantial strength decrease of CC samples in the range of 300 and 400 °C was

reported to be due to the dissociation of calcium hydroxide while the strength loss

between 500 and 600 caused by dehydration of calcium hydroxide (Mendes,

Sanjayan et al. 2008). At 800 °C, HGC and OGC samples displayed higher strength

of about 44% and 38% of their initial strength, respectively than 26% for CC

samples.

Page 134: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

119

Figure 5.11 Comparison of the reduction factors of compressive strength for OGC, HGC

and CC at various elevated temperatures

The strain corresponding to the peak strength values for OGC, HGC and CC at

temperatures of 23, 200, 400, 600 and 800 °C tested in load control regime are

shown in Figure 5.12. The strain associated with peak strength did not substantially

change for HGC and CC mixes up to 600 °C followed by an increase up to 800 °C.

On the other hand, OGC mixes showed moderately increasing trend of strain up to

400 °C followed by a significant increase above this temperature showing a

significant change in the material behaviour from solid to viscoelastic. The strains

attributed to the peak strength at 800 °C were 4.40, 2.93 and 2.68 times the strain at

room temperature for OGC, HGC and CC, respectively. This shows that geopolymer

sustained large deformations before failure. This behaviour of geopolymers (large

strain) is similar to the behaviour of glasses and metals at high temperatures due to

their atomic or molecular self diffusion mechanism (Pan and Sanjayan 2010).

0

0.2

0.4

0.6

0.8

1

1.2

0 200 400 600 800 1000

Red

uct

ion

fac

tor

of

stre

ngt

h

Temperature (° C )

OGC HGC CC

Page 135: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

120

Figure 5.12 Strain at peak stress for OGC, HGC and CC at various elevated Temperatures

The changes in the physical appearance of the OGC, HGC and CC samples are

illustrated in Figures 5.13 to 5.15, respectively. As it can be seen there was no

significant change of colour in OGC and HGC samples up to 400 °C while both

showed very similar surface colour of slightly light brown. Both samples displayed

medium brown at 600 °C and clearly dark brown at 800 °C. On the other hand, CC

samples exposed to elevated temperatures as shown in Figure 5.15 displayed

negligible change in terms of colour. The visible change of colour in OGC and HGC

samples might be associated with higher content of iron oxide in fly ash and their

oxidisation at elevated temperatures (Kong and Sanjayan 2010).

In many of the samples, surface cracking was observed which initiated from

temperature differential of the surface and the centre of cylinders. Also they might be

0

0.002

0.004

0.006

0.008

0.01

0.012

0 100 200 300 400 500 600 700 800 900

Stra

in a

t p

eak

stre

ss

Temperature (º C)

OGC HGC CC

Page 136: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

121

associated with differential strain caused by temperature gradient through concrete

cross section (Sarker, Kelly et al. 2014).

Figure 5.13 Effect of elevated temperatures on the physical appearance of one-part

geopolymer concrete (OGC)

Figure 5.14 Effect of elevated temperatures on the physical appearance of one-part hybrid

OPC-geopolymer concrete (HGC)

Page 137: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

122

Figure 5.15 Effect of elevated temperatures on the physical appearance of OPC concrete

(CC)

Thermal properties

The measured values of thermal conductivity for OGC, HGC and CC mixes

subjected to the temperature range of 23 to 600°C are compared in Figure 5.16. In

general, at all temperatures, CC indicated the highest values of thermal conductivity

and the lowest results were achieved by OGC. At room temperature, the thermal

conductivity of CC, HGC and OGC were 1.717, 1.418 and 1.089 W/mK,

respectively. The results suggest that by replacing 70 % of OPC with fly ash in HGC

samples, the thermal conductivity decreased by 17%. The samples containing 50

wt% fly ash and 50 wt% slag (OGC) exhibited a further 23% decrease. This

behaviour was also reported in the literature that by increasing the content of fly ash,

thermal conductivity was significantly decreased. The lower thermal conductivity of

Al-Si systems results from the lower moisture content compared to OPC mixes while

these systems have different pore structures (Pan, Tao et al. 2018). The general trend

Page 138: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

123

of thermal conductivity development by increasing temperature of the OGC mixes

was different from those of HGC and CC samples. The overall trend of CC and HGC

systems was that thermal conductivity decreased with increasing temperature up to

400 °C and then remained relatively stable above this temperature. For OGC mixes,

the thermal conductivity dropped up to 200 °C and beyond this range became almost

constant. The distinct behaviour of mixes influenced by increasing temperature

exhibits the different physiochemical reactions that occur for each mix. For the CC

systems, the rapid decrease of thermal conductivity up to 400 °C was attributed to

the moisture loss while the minor changes above this range were associated with the

changes of moisture content and porosity of the CC system (Khaliq and Kodur

2011). For OGC mixes, the monotonic behaviour in terms of thermal conductivity

might be associated with densification of the pore structure with increasing

temperature.

Figure 5.16 Effect of temperature on thermal conductivity

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

0 100 200 300 400 500 600

Ther

mal

co

nd

uct

ivit

y (

W/m

K )

Temperature (ºC)

OGC HGC CC

Page 139: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

124

The measured values of specific heat are exhibited in Figure 5.17. The CC mix

achieved the highest value of 2.614 MJ/m3K at room temperature in comparison with

2.03 and 1.91 MJ/m3K for OGC and HGC mixes, respectively. All three mixes

showed similar increasing pattern up to 100 °C. This increase in specific heat was

due to the physical densification effect resulting from drying which in turn caused

reduction of permeability (Pan, Tao et al. 2018). For CC mix, the specific heat

remained stable in the range of 100 and 400 °C followed by a significant increase at

500 °C. This higher specific heat value at 500 °C was associated with the absorption

of heat in the hydration reaction (Kodur and Khaliq 2010). The specific heat of HGC

mixes dropped significantly in the range of 100 and 300 °C which is due to the

conversion of moisture into vapour and was also observed in the previous findings

for alkali activated slag binders. This decrease was associated with a higher specific

heat of original products compared with the converted products while the final

product required extra energy for bending and breaking of the hydrogen bonds

(Ukrainczyk and Matusinović 2010, Pan, Tao et al. 2018). The OGC samples

experienced decrease of specific heat from 200 to 300 °C. This exothermic peak

might be due to conversion of moisture into vapour as well. The specific heat then

increased up to 600 °C.

Page 140: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

125

Figure 5.17 Effect of temperature on specific heat

Evaluation of Test Results with Code Predictions

Modulus of elasticity (MOE)

There are a number of predictions on elastic modulus of concrete which depending

on the paste and type of aggregate. The measured values for modulus of elasticity of

OGC, HGC and CC are compared with some of the proposed empirical equations for

OPC concrete (Equations 5.4-5.6) and for geopolymer concrete (Equation 5.7) as

follows:

American Concrete Institute, ACI 363 (1993) :

𝐸𝑐 = 3320√𝑓𝑐′ + 6900 (5.4)

Australian Standards, AS 3600 (2009):

𝐸𝑐 = 0.043𝜌1.5√𝑓𝑐′ (5.5)

1

1.5

2

2.5

3

3.5

4

4.5

0 100 200 300 400 500 600

Spec

ific

hea

t (

MJ/

m3

K)

Temperature (ºC)

OGC HGC CC

Page 141: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

126

Ahmad and Shah (Ahmad and Shah 1985)

𝐸𝑐 = 3.38𝜌2.5(√𝑓𝑐′)0.65 × 10−5 (5.6)

Hardjito et al. (Hardjito, Wallah et al. 2005)

𝐸𝑐 = 2707√𝑓𝑐′ + 5300 (5.7)

The trend lines through the predicted values of Equations 5.4 to 5.7 and the measured

values for OGC, HGC and CC are shown in Figure 5.18. The modulus of elasticity of

OGC, HGC and CC could be predicted reasonably well by using Equation 5.7,

Equation 5.4 and Equation 5.6, respectively.

Figure 5.18 Measured MOE against predictions

Page 142: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

127

Stress-strain relationship of concrete

The number of research on the stress-strain curves of geopolymer concrete is very

limited. The stress-strain models of Popovics (Popovics 1973) modified by

Thorenfeldt et al. (Thorenfeldt 1987) for OPC concrete is used in this study to

evaluate its suitability for the developed OGC and HGC mixes by the following

expression:

𝑓𝑐

𝑓𝑐′ =

𝜀𝑐

𝜀𝑐′ .

𝑛

𝑛−1+(𝜀𝑐

𝜀𝑐′ )𝑛𝑘

(5.8)

Where fc = concrete compressive stress, 𝜀𝑐 = strain in concrete, 𝑓𝑐′= maximum

compressive stress in concrete, 𝜀𝑐′ = strain where 𝑓𝑐 reaches 𝑓𝑐

′ and n = curve fitting

factor. The factor k is equals 1 when 𝜀𝑐

𝜀𝑐′ is less than 1 and for greater than 1 , this

value is measured using Equation 5.9 proposed by Collins and Mitchell (Collins and

Mitchell 1991):

𝑘 = 0.67 +𝑓𝑐

62 (in MPa unit) (5.9)

The value of strain at peak stress (𝜀𝑐′ ) could be found using Equation 5.10 suggested

by Collins et al. (Collins, Mitchell et al. 1993):

𝜀𝑐′ =

𝑓𝑐′

𝐸𝑐.

𝑛

𝑛−1 (5.10)

Two different Equations of 5.11 suggested by Collins and Mitchell (Collins and

Mitchell 1991) and 5.12 suggested by Sarker (Sarker 2009) were used to measure

the curve fitting factor n. The values calculated by Equation 5.12 were proved to

provide a better estimation of the curve fitting factor for the fly ash-based

Page 143: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

128

geopolymer concrete. The stress-strain curves calculated by using the Equations 5.11

and 5.12 are compared with the tested results as shown In Figure 5.19.

𝑛 = 0.8 +𝑓𝑐

17 (in MPa unit) (5.11)

𝑛 = 0.8 +𝑓𝑐

12 (in MPa unit) (5.12)

Figure 5.19 Stress-strain curves of OGC, HGC and CC versus curves calculated by

Equation 5.11 (Thorenfeldt 1987) and Equation 5.12 (Sarker 2009)

From the comparison of tested and measured stress-strain curves, it can be

understood that Equation 5.12 provides a better estimation of the curve fitting factor

for OGC, while Equation 5.11 provides more suitable prediction in case of HGC and

CC.

Page 144: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

129

Compressive reduction factors at elevated temperatures

In order to examine the influence of aggregate type on the reduction factors of the

compressive strength at elevated temperatures, the test results are measured with the

proposed reduction factors in Eurocode 2 (2005) and ASCE (1992) model as shown

in Figure 5.20.

Figure 5.20 Comparison of the tested strength reduction factor with existing models

It is notable that the results of the tested strength reduction factors in the range of 100

℃ and 300 ℃ were lower than both Eurocode and ASCE models which was also

reported in previous studies as a result of water evaporation and transformation

(Potha Raju, Srinivasa Rao et al. 2007, Michikoshi 2008, Tao, Yuan et al. 2010).

However, beyond this range, both models underestimated the strength for OGC and

HGC and use of these models for predicting the strength at elevated temperatures is

very conservative.

0

0.2

0.4

0.6

0.8

1

1.2

0 200 400 600 800 1000

Red

uct

ion

fac

tor

Temperature (° C )

OGC

HGC

CC

Euro-Sili

Euro-Car

ASCE

Page 145: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

130

Thermal conductivity

The comparison of the measured thermal conductivity of OGC, HGC and CC with

Eurocode 2 (2005) models is illustrated in Figure 5.21. As can be seen, the measured

thermal conductivity of CC and HGC had acceptable accordance with the upper and

lower limits of Eurocode, while the measured thermal conductivity for OGC did not

locate within the upper and lower limit in the range of 20 ℃ to 500 ℃. Thus the

Eurocode predications cannot be used for this type of concrete and further research is

required to develop empirical models that could be used to estimate the thermal

properties of OGC at elevated temperatures.

Figure 5.21 Comparison of the measured thermal conductivity with existing model

Conclusions

Different material properties of ambient cured one-part hybrid OPC-geopolymer

concrete (HGC), one-part geopolymer concrete (OGC) and conventional OPC

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

2.2

0 100 200 300 400 500 600

Ther

mal

co

nd

uct

ivit

y (

W/m

K )

Temperature (ºC)

OGC

HGC

CC

Euro-upper

Euro-lower

Page 146: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

131

concrete (CC) have been evaluated in the laboratory including mechanical and

thermal properties. To conclude this chapter, the important findings of the chapter

are presented as follows:

• Although OGC, HGC and CC mixes were composed of the same amount of

aggregate and binder and also similar water to binder ratio, they exhibited

different mechanical properties. The 28 day compressive strength of OGC

and HGC were in the range of 40 MPa compared to 50 MPa for CC.

Furthermore, CC mixes achieved higher MOE, MOR and indirect tensile

strength compared to OGC and HGC.

• OGC, HGC and CC mixes showed different behaviour in terms of stress-

strain relationship curves. HGC and CC samples exhibited increase in the

initial slope of the curve and more ductile post-peak behaviour while OGC

failed suddenly and showed the least ductility among others.

• In terms of compressive strength of concrete mixes at various elevated

temperatures, OGC, HGC and CC samples exhibited similar behaviour in the

range of 23 to 200 °C but different behaviour beyond this range. The decrease

in compressive strength between 23 to 200 °C was associated with a number

of temperature-induced stresses and chemical reactions and mainly results

from different thermal incompatibilities between the paste and aggregate.

OGC mixes exhibited 32% increase of strength up to 600 °C while HGC

samples experienced 36% increase up to 400 °C followed by a decrease

above this temperature. On the other hand CC samples exhibited an overall

strength decline trend in all temperatures. At 800 °C, HGC and OGC samples

Page 147: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

132

displayed higher strength of about 44% and 38% of their initial strength,

respectively compared to 26% for CC samples.

• The effect of increasing temperature on the strain at peak strength of OGC,

HGC and CC mixes were also investigated. In this regard, HGC and CC

samples showed almost similar patterns while the strain did not change

considerably up to 600 °C followed by an increase between 600 and 800 °C.

On the other hand, OGC mixes showed moderately increasing trend of strain

up to 400 °C followed by a significant increase above this temperature,

showing a significant change in the material behaviour. The strains attributed

to the peak strength at 800 °C were 4.40, 2.93 and 2.68 times the strain at

room temperature for OGC, HGC and CC, respectively.

• OGC, HGC and CC samples exhibited different behaviour in terms of

measured thermal conductivity and specific heat values by increasing

temperature which is associated with their different gel structures. At elevated

temperatures, the thermal conductivity of HGC and CC samples decreased up

to 400 °C and then remained almost stable while for OGC mixes, the thermal

conductivity slightly decreased between 23 and 200 °C and was constant

beyond this range. Generally, the specific heat increased by increasing

temperature but a significant decrease was observed for OGC and HGC

associated with conversion of moisture into vapour.

• Comparison of the measured modulus of elasticity with the existing models

showed that the prediction of the modulus of elasticity suggested by Hardjito

et al. (Hardjito, Wallah et al. 2005) was close to OGC, ACI 363 (1993) was

close to HGC and Ahmad and Shah (Ahmad and Shah 1985) had good

Page 148: Mechanical, Thermal and Structural Investigations of One

Chapter 5 Mechanical and Thermal Properties

133

agreement with CC test results of this experiment. From the comparison of

tested and measured stress-strain curves, it was concluded that Equation 5.12

provides a better estimation of the curve fitting factor for OGC, while

Equation 5.11 provides more suitable prediction in case of HGC and CC.

In terms of compressive strength at elevated temperatures, it was observed that the

results of the tested strength reduction factors in the range of 100 ℃ and 300 ℃ were

lower than both Eurocode and ASCE models. However, beyond this range, both

models underestimated the strength for OGC and HGC and use of these models for

predicting the strength at elevated temperatures is very conservative. Also, the

measured thermal conductivity of CC and HGC had acceptable accordance with the

upper and lower limits of Eurocode, while the measured thermal conductivity for

OGC did not locate within the upper and lower limit in the range of 20℃ to 500℃.

Thus the Eurocode predications cannot be used to predict the thermal conductivity of

this type of concrete.

Page 149: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

134

Chapter 6 Performance of Reinforced Concrete

Columns at Ambient Temperature

General

In Chapter 5, the basic mechanical properties of different types of concrete

including one-part hybrid OPC-geopolymer concrete (HGC), one-part geopolymer

concrete (OGC) and ordinary Portland cement concrete (CC) were investigated. The

feasibility of using HGC and OGC to replace CC has been demonstrated. Despite the

fact that using HGC and OGC can achieve environmental benefits as well as

reasonable mechanical properties, the potential applications of these new types of

concrete have not been well explored. Compared with the extensive studies

conducted on geopolymer concrete material, very limited research has been carried

out on structural members with geopolymer concrete as reviewed in Chapter 2. Lack

of a suitable analytical model may be one of the reasons to limit the application of

geopolymer concrete in structures. In this chapter, the effects of different types of

concrete including HGC, OGC and CC on the performance of reinforced columns at

ambient temperatures will be investigated. The geometry of specimens, material

properties, a detailed description of the test program and experimental results will be

described. Two columns made of HGC and OGC as new types of concrete were

tested in ambient condition. The effects of concrete type on the failure modes,

development of axial and lateral strains and strength of reinforced columns is

discussed. In addition, the test results are also compared with predictions of finite

element (FE) analysis for HGC and OGC columns compared with CC columns.

Page 150: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

135

Experimental Program

Specimen preparation and material properties

A total of 2 square columns, including 1 reinforced HGC and 1 reinforced OGC were

fabricated. The columns were designed according to AS3600 (2009) and were 1540

mm long and of square cross section of 200 mm. The same reinforcing bars of 16

mm were used to build steel cages. The bars were tied with 10 mm ties at a spacing

of 150 mm the whole length. The details of columns elevation and cross section are

depicted in Figure 6.1.

Figure 6.1 Elevation and cross section of the columns

Page 151: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

136

After assembling steel cages, they were placed into the plywood moulds and uniform

concrete cover was provided in all directions using plastic chairs and wheel spacers

(see Figure 6.2).

The material properties of longitudinal bars were determined by tensile tests at

ambient temperature according to AS-1391 (2007). Three tensile coupons for

reinforcing steel were tested. A typical stress-strain curve obtained from the tests is

shown in Figure 6.3 and the average material properties are listed in Table 6.1, where

E0 is the modulus of elasticity at ambient temperature; fy is the 0.2 % proof strength;

n is the strain hardening exponent; fu is the ultimate strength; 𝜀𝑢 is the strain

corresponding to ultimate strength. For the Ф10 stirrup, the nominal yield strength

was 500 MPa.

Figure 6.2 Preparation of RC columns

Page 152: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

137

Three different types of concrete, including OGC, HGC and CC, with mix

proportions summarised in Table 5.2 were used to fill the columns. The mechanical

properties of concrete mixes at 28 days of age were presented in Table 5.3 and the

measured cylinder compressive strength 𝑓𝑐′ on the testing day are shown in Table 6.2.

Due to the limitation of vibrator, the columns were cast horizontally. After casting,

the concrete was compacted and the surface was levelled. Then, the columns were

covered by plastic sheet and were stored in the laboratory until testing. The

appearance of the fabricated reinforced columns is shown in Figure 6.4, indicating

fairly good quality of fabrication.

Table 6.1 Material properties of reinforced steel

Dimeter

(mm)

𝐸0

(MPa)

𝑓𝑦

(MPa)

𝑓𝑢

(MPa)

𝜀𝑢

(%)

12 197800 528.1 631.2 7.8

Figure 6.3 Stress-strain curve for reinforcing steel

0

200

400

600

800

0 5 10 15 20

Stre

ss (

MP

a)

Strain ε %

Page 153: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

138

Figure 6.4 Fabricated RC columns, (a) after casting, (b) at testing day

Test procedure and setup

The experiments were performed in the Structure Laboratory at the University of

Western Sydney, Australia. The specimens were tested under axial compression

using the testing machine with a maximum capacity of 2,000 kN as shown in Figure

6.5. Two sets of end plates were attached to the top and bottom ends of the columns

to apply the load to the column at axial centric. The test procedure proposed by Tao

and Yu (Tao and Yu 2008) was accepted for this experiment. Several linear variable

displacement transducers (LVDTs) were attached to the surface of column before

testing. For each test specimen, four LVDTs with a range of 100 mm were installed

in the mid-height region including two for measuring the axial strains and two for

lateral strains. Two LVDTs at 180º apart at the top and two at the bottom were used

to measure the axial shortening during the tests and two LVDTs were used at quarter

span of the column to monitor the deflection. The detailed locations of the LVDTs

are shown in Figure 6.5.

Axial loading was applied with displacement control at a rate of 0.5 mm/min. The

eccentricity of each column was measured before applying the compression load and

(a) (b)

Page 154: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

139

the results are shown in Table 6.2. The effective length of the column Le, defined as

the distance between the centre of top hinge and bottom hinge and was 1970 mm for

all test specimens, leading to a slenderness ratio (Le/r) of 33. According to AS 3600

(2009), these columns are categorised as slender columns.

During the elastic range of the loading procedure, the specimen was loaded at a rate

of approximately 2 kN/s. The load interval was 1/10 𝑁𝑢 and was reduced by 50%

upon reaching 0.75 𝑁𝑢 (𝑁𝑢: predicted ultimate capacity of columns). At each load

interval the load was sustained for 1 min to record the readings after they were

stable. At higher load levels, the axial compressive load was applied continually to

capture the maximum compressive load of the specimen. The tests terminated when

the axial displacement reached 20 mm.

It should be mentioned that due to the testing machine capacity limit, the

compression testing up to failure for CC column was not achievable and accordingly,

the experiment for this type of concrete was not carried out. Also, due to the

abundance of existing guidelines and models for prediction of reinforced OPC

concrete columns behaviour, a FE model was developed for CC column and the

results were compared with those of OGC and HGC columns.

Page 155: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

140

Figure 6.5 Test set-up

Table 6.2 Details of reinforced columns

Concrete

type

𝒇𝒄′ on the testing day

(MPa)

Eccentricity at the top

(eTop) (mm)

Eccentricity at the bottom

(eBottom) (mm)

OGC 41.3 8.1 13.4

HGC 46.6 5.5 2.8

Page 156: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

141

Experimental Results and Discussion

General observations and failure modes

The failure appearances of the 2 reinforced OGC and HGC columns are presented in

Figures 6.6 and 6.7. By increasing the compressive load, cracks initiated on the

tension faces of both columns and propagated by further increase of load. Both

columns suddenly failed in a brittle and explosive manner, characterised by the

compressive crushing of concrete with a short post-peak behaviour. The failure

regions were located within the bottom ends and mi-height of the specimens. For

both specimens, spalling of the concrete cover decreased the load carrying capacity

and accordingly overall instability and buckling of the longitudinal reinforcement

bars was observed.

Figure 6.6 Failure mode of OGC column

(a) (d) (c) (b)

Page 157: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

142

Figure 6.7 Failure mode of HGC column

Axial load versus axial displacement curves

The effect of concrete type on the axial load N versus axial displacement curve is

depicted in Figure 6.8. The displacement values were obtained from the

measurements of the LVDTs. For both columns, the initial buckling occurred at the

peak load or just before the peak load. From Figure 6.8 it can be seen that the axial

displacement value for OGC is 29% higher than the one for HGC column at peak

load. This improvement could be due to the delay of cover spalling in OGC column

compared to HGC column.

The capacity of reinforced concrete columns are mainly contributed by the strength

of concrete. Also, the load eccentricity and specimen imperfections have some

influence on the ultimate capacity of columns. The use of HGC with 12% higher

compressive strength compared to OGC, resulted in an increase of 37% of ultimate

strength for reinforced concrete column. It should be mentioned that the load for

OGC column was applied with higher eccentricities compared to HGC specimen as

(a) (d) (c) (b)

Page 158: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

143

presented in Table 6.3, which might be another reason for further decrease of

strength for this column. To make this clear, a sensitivity analysis will be performed

in section 6.5 using FE model to justify the variation of ultimate capacity among

reinforced columns.

It should also be noted that due to the sudden failure of columns, the descending

branch of the graph could not be captured.

Figure 6.8 Axial displacement versus axial load

Lateral deformation

The load versus the mid-height deflection behaviour of OGC and HGC columns are

shown in Figure 6.9. Mid-height lateral displacements were measured by LVDTs

installed at column mid-height. For both columns, an initial linear branch developed

up to a load level of approximately 20% for OGC and 25% for HGC column. The

initial lateral stiffness was slightly higher for OGC column compared to that of HGC

column. After this stage, a semi-linear ascending branch continued to peak load.

0

200

400

600

800

1000

1200

1400

1600

1800

2000

0 1 2 3 4 5 6 7 8

Axi

al lo

ad N

: K

N

Axial dispalcement (mm)

OGC HGC

Page 159: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

144

After reaching the peak load, the displacement significantly increased. At the peak

load, the mid-height displacement for OGC column was 0.9 versus 2.1 mm for HGC

column. Therefore, the use of HGC in reinforced columns, increased the capacity and

lateral deflection of the column compared to OGC.

Figure 6.9 Mid-height deflection versus axial load

Evaluation of Test Results with FE Predictions

Materials modelling

There has been vast number of FE modellings to predict the behaviour of reinforced

columns with normal concrete (Claeson and Gylltoft 1998, Nmecek and Bittnar

2004, Němeček, Padevět et al. 2005, Majewski, Bobinski et al. 2008). More recently,

Sarker (Sarker 2009) employed a non-linear FE analysis to evaluate the suitability of

using the modified Popovics stress-strain model to predict the behaviour of tested

slender geopolymer concrete columns. The analytical model reported to provide

good accuracy to predict the ultimate load and mid-height deflection of the columns.

0

200

400

600

800

1000

1200

1400

1600

1800

2000

0 0.5 1 1.5 2 2.5

Axi

al lo

ad N

: kN

Mid-height deflection

OGC HGC

Page 160: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

145

ABAQUS software (2012) has used in this study to develop the FE model. The

concrete damage plasticity (CDP) model (Lee and Fenves 1998) in ABAQUS was

used for modelling the concrete crack behaviour. In the CDP model, compression

and tension behaviour are defined by introducing two different damage parameters

into a plasticity-based model. It should be mentioned that the mechanical properties

relationships of OGC, HGC and CC presented in Chapter 5 are used for FE model.

For the behaviour of concrete in tension, the material model of Nayal et al. (Nayal

and Rasheed 2006) was used. For stress-strain relationship of concrete, the Popovics

model modified by Thorenfeldt et al. (Thorenfeldt 1987) with different curve fitting

factors as discussed in Chapter 5 was used.

A flow potential eccentricity of e = 0.1 and dilation angle of = 38○ as

recommended by Jankowiak et. al. (Jankowiak and Lodygowski 2005) were adopted

for the plasticity parameters and a biaxial to uniaxial compressive strength ratio of

1.16 was used as per the ABAQUS (2012) user manual recommendation. The

stiffness degradation after crushing (compression) and cracking (tension) of concrete

was considered using the two separate damage laws of Dc= 1- fcm /c and Dt = 1- fct

/t , respectively. The stress-strain curves in tension and compression and the

evolution of damage with the inelastic strains are shown in Figure 6.10.

Page 161: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

146

Figure 6.10 Adopted material behaviour for cementitious grout, (a) stress-strain under

compression, (b) evolution of compressive damage index Dc versus inelastic strain, (b)

stress-strain under tension and (d) evolution of tensile damage index Dt versus cracking

strain.

For the steel material behaviour, an elastic-isotropic plastic hardening material model

was used. The mechanical properties of reinforcing steel were presented in Table 6.1

and the stress-strain curve was shown in Figure 6.3. The Possion’s ratio for steel

components were taken as ν = 0.3.

Other details of FE model

8-node brick elements with three translation degrees of freedom at each node

(C3D8R) and 2-node Beam elements were used to model the concrete and

reinforcement steel, respectively. For optimal FE mesh, an element size of 12 mm

was chosen.

Page 162: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

147

A tie constraint was used to simulate the interaction of steel and concrete, in which

the degree of freedom on beam and brick elements were tied together. Slip between

reinforcements and concrete has been neglected due to simplicity. Also, a surface-to-

surface contact was used to simulate the interaction of the column and steel

endplates, in which “Hard contact” was selected to specify the normal direction and

“Coulomb friction: with a coefficient of friction as 0.3 was adopted for tangent

contact simulation. In the model, the top and bottom surfaces of the column can be

fixed against all degrees of freedom except for the displacement at the loaded end.

The FE model was analysed using a displacement-control procedure in a quasi-static

state. A maximum prescribed displacement was applied on the top plate.

Comparison of predictions with experimental results

The mentioned model in ABAQUS was used to perform the column analysis. The

predicted axial load versus mid-height deflection and axial load versus axial

displacement for OGC and HGC columns are compared with the test results in

Figures 6.11 and 6.12, respectively and the results are also summarised in Table 6.4.

In general, the proposed FE model provided reasonable predictions for both OGC

and HGC columns. The FE model provided very close estimation of the ultimate

strength of reinforced OGC and HGC columns with a mean value of the test-

prediction of 1.02 with a standard deviation of 6.3%. The FE slightly underestimated

the ultimate strength of OGC columns (PFE/Ptest=0.93) and marginally overestimated

the ultimate strength of the HGC columns (PFE/Ptest=1.02). It can be seen that the FE

predictions are almost on the safe side for the reinforced OGC and HGC columns.

Page 163: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

148

Figure 6.11 Comparison between measured and predicted Load-Deflection curve for OGC

and HGC columns

Figure 6.12 Comparison between measured and predicted Load-Axial displacement curve

for OGC and HGC columns

The ultimate capacity of the CC reinforced column and its axial displacement and

mid-height deflection at peak load are also calculated using the developed FE model

and the results are shown in Figure 6.13 and Table 6.3.

Page 164: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

149

Figure 6.13 Predicted Load-Deflection and Load-Axial displacement curves for CC columns

Table 6.3 Summary of the experimental test results and numerical modelling

Column

Ultimate

Capacity (kN)

Ptest/PFE Axial

Displacement (mm)

Mid-height

Deflection (mm)

OGC Ptest 1402

1.066

6.2 0.9

PFE 1315 6.3 1.0

HGC Ptest 1924

0.977

4.4 2.1

PFE 1968 3.4 1.8

CC PFE 2470 - 3.4 1.8

Average 1.0215

Standard deviation 0.063

Page 165: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

150

Figure 6.14 Typical FE analysis (a) tension damage (b) compression damage (c) maximum

stress on concrete (d) maximum stress on reinforcement

Sensitivity analysis

A sensitivity analysis was carried out to identify the important factors in behaviour of

reinforced columns. The effects of variation of compressive strength of OGC and

HGC and load eccentricities on the ultimate capacity of the reinforced concrete

columns were evaluated. Details of parameters and results of sensitivity analysis are

depicted in Table 6.4. The results show that both concrete compressive strength and

eccentricity have substantial influence on the ultimate capacity of reinforced

columns. However, variation of concrete strength has more influence on the ultimate

load capacity of reinforced concrete columns. For instance, by increasing 43% of

compressive strength of OGC, around 44% increase in load capacity was observed,

while by increasing 50% load eccentricity, 15% decrease of capacity was observed.

Page 166: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

151

The reasons for the lower capacity of OGC column in comparison with HGC column

could be justified by the results of this analysis, which are the lower compressive

strength of OGC and higher load eccentricities for this column.

Table 6.4 Summary of sensitivity analysis parameters and results

Concrete Type

Concrete

compressive

strength (MPa)

Load eccentricity

ratio (e/D)

Failure load

(kN)

OGC

35.0 0.05 1372.9

41.3 0.05 1642.0

45.0 0.05 1811.5

50.0 0.05 1975.5

41.3 0.08 1510.9

41.3 0.12 1313.2

HGC

37.0 0.05 1492.2

46.6 0.05 1843.62

52.0 0.05 2034.2

46.6 0.08 1677.7

46.6 0.12 1454.4

Verification of current test results

The developed model was also used to predict the behaviour of the slender

geopolymer concrete columns tested by Sumajouw et al. (Sumajouw, Hardjito et al.

2007). For this purpose, two columns of OGCI-1 and OGCI-4 are selected. The

column cross section was a 175 mm square with a length of 1500 mm. The average

compressive strength of concrete was 40 MPa for these two selected columns. The

longitudinal reinforcement was four 12 mm (N 12) deformed bars for OGCI-1 and

eight 12 mm (N12) deformed bars for OGCI-4 column with similar lateral

reinforcement of 6 mm. The compressive load was applied with 15 mm eccentricity

for both columns. The comparison of the FE results with test results reported by

Page 167: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

152

Sumajouw et al. for axial load versus mid-height deflection of OGCI-1 and OGCI-4

are depicted in Figures 6.15 and 6.16, respectively and the results are also

summarised in Table 6.5. It can be seen that the FE analysis gives very close

predictions for the failure load of the geopolymer slender columns. The load capacity

of the columns correlated very well with the value calculated using the FE model

with a mean value of the test-prediction of 1.002 with a standard deviation of 9.8%.

Figure 6.15 Axial load versus mid-height deflection curves for OGCI-1

0

500

1000

1500

0 1 2 3 4 5 6 7 8

Load

(kN

)

Deflection (mm)

FE

Test by Sumajouw et al. (Sumajouw, Hardjito et al. 2007)

Page 168: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

153

Figure 6.16 Axial load versus mid-height deflection curves for OGCI-4

Table 6.5 Summary of the experimental results by Sumajouw et al. (Sumajouw, Hardjito et

al. 2007) and numerical modelling of OGCI-1 and OGCI-4 specimens

Column Stiffness

(kN/mm)

Ultimate

Capacity (kN)

Exp/FE Mid-height

Deflection (mm)

OGCI-1 Exp. 318 940

0.932

5.4

FE 358 1008 4.8

OGCI-4 Exp. 386 1237

1.072

6.2

FE 425 1153 4.5

0

500

1000

1500

0 1 2 3 4 5 6 7 8

Load

(kN

)

Deflection (mm)

FE

Test by Sumajouw et al. (Sumajouw, Hardjito et al. 2007)

Average 1.002

Standard deviation 0.098

Page 169: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

154

Conclusions

The behaviour of axially loaded reinforced concrete columns with three different

types of concrete including OGC, HGC and CC were experimentally and

numerically investigated in this chapter. To conclude this chapter, the important

findings are presented as follows:

• The experimental results indicated that the compressive strength of concrete

and load eccentricity had significant influence on the ultimate capacity as

well as mid-height deflection and axial strain at peak load. The axial

displacement value for OGC was 29% higher than the one for HGC column

at peak load. However, the initial stiffness in OGC column was lower than

HGC column. At peak load, the mid-height displacement for OGC column

was 0.9 versus 2.1 mm for HGC column. Therefore, the use of HGC in

reinforced columns with higher compressive strength, increased the ultimate

capacity and lateral deflection of the column compared to OGC.

• The proposed FE model in this study, can predict the test results of

reinforced columns with different types of concrete such as OGC and HGC

reasonably well.

The results presented in this chapter demonstrate that one-part geopolymer concrete

(OGC) and one-part hybrid OPC-geopolymer concrete (HGC) have excellent

potential for application in the building industry while they provide comparable

initial stiffness, strength and ductility compare to conventional concrete OPC (CC)

columns. Further research is required to consider the influence of different

parameters such as load eccentricity, imperfections, strength, longitudinal

reinforcement and transverse reinforcement etc on the behaviour of reinforced OGC

Page 170: Mechanical, Thermal and Structural Investigations of One

Chapter 6 Performance of Reinforced Concrete Columns at Ambient Temperature

155

and HGC columns to modify the existing design provisions for these types of

concrete.

Page 171: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

156

Chapter 7 Performance of Reinforced Concrete

Columns at Elevated Temperatures

General

Fire safety has become a major concern in the engineering design of building

structures. As discussed in the literature review in Chapter 2, and based on the

experimental results of Chapter 5, geopolymer concrete has the potential to improve

the fire performance of structural members due to its excellent thermal and

mechanical properties. However, research works on the fire performance of

geopolymer concrete members are very limited. Thus, this chapter introduces a series

of investigations on the behaviour of reinforced one-part hybrid OPC-geopolymer

concrete (HGC) and one-part geopolymer concrete (OGC) columns in comparison

with ordinary Portland cement concrete (CC) columns. Ten reinforced concrete

columns were tested in fire condition at constant load level. Three of these columns

were made of OGC, three were made of HGC and four were made of CC. The

influencing parameters were studied including, concrete type (OGC, HGC and CC)

and load levels (0.3-0.45 of ultimate). The geometry of specimens, the location of

thermocouples, material properties, a detailed description of the test program and

experimental results will be described in this chapter. The aim of this test program

was to address the gap in the research on the behaviour of geopolymer concrete

columns in fire condition. The effects of concrete type and load level on the failure

modes, fire resistance and axial strains are discussed. The experimental results are

then compared with predictions of the developed FE model.

Page 172: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

157

Experimental Program

Specimen preparation and material properties

A total of 10 square columns, including 3 reinforced OGC, 3 reinforced OGC and 4

reinforced CC columns were fabricated. The columns were designed according to

AS3600 (2009) and were 1,540 mm long including two steel endplates, each with a

thickness of 20 mm and of square cross section of 200 mm. The calculated length

from the centre of the upper hinge to the centre of the lower hinge is 1,970 mm. The

same reinforcing bars of 16 mm were used to build steel cages. The bars were tied

with 10 mm ties at a spacing of 150 mm over the whole length. The details of

columns elevation and cross section are depicted in Figure 7.1. Three thermocouples

were embedded into the mid-height section of each specimen before concrete was

poured. One thermocouple was located at one-quarter of the column depth, one was

embedded in the centre and one was attached to one of the longitudinal reinforcing

bars. Also, one thermocouple was attached to the surface of the column. The

locations of thermocouples are depicted in the cross-section in Figure 7.1.

The following parameters were evaluated in this test:

(1) Concrete type: OGC, HGC and CC.

(2) Load level (𝑛𝑓): in the range of 0.2 and 0.45 of ultimate, where 𝑛𝑓 = 𝑁0/𝑁𝑢

while 𝑁0 is the applied axial compression load and 𝑁𝑢 is the ultimate bearing

capacity of the column at ambient temperature. 𝑁𝑢 was measured by testing

of reinforced OGC and HGC columns at ambient temperature and was

calculated through FE analysis for CC column and the results were presented

in Chapter 6.

Page 173: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

158

The mix proportions of different types of concrete were summarised in Table 5.2 and

their mechanical properties were presented in Table 5.3. The properties of OGC,

HGC and CC cylinders subjected to elevated temperatures and their thermal

properties were also investigated in Chapter 5. The mechanical properties of

reinforcing steel were presented in Table 6.1 and the stress-strain curve was shown in

Figure 6.3. Furthermore, the details of columns assembly and fabrication were

discussed in Chapter 6. The details of the columns specimens tested in fire and the

measured eccentricity for each column are given in Table 7.1.

Figure 7.1 Detailed dimensions of the reinforced column specimens

Page 174: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

159

Table 7.1 Details of concrete columns tested in fire

Specimen

label

Age at

testing day

𝒇𝒄′ on the

testing day

(MPa)

Load level

nf

Applied load

N0 (kN)

Top

eccentricity

eTop (mm)

Bottom

eccentricity

ebottom (mm)

OGC-1a 160 42.3 0.33 462 7.84 5.10

OGC-1b 182 42.8 0.33 462 7.17 10.01

OGC-2 175 41.8 0.45 630 3.67 3.48

HGC-1a 162 44.5 0.33 635 3.72 3.05

HGC-1b 188 45.6 0.33 635 2.18 1.92

HGC-2 177 44.5 0.45 860 7.12 9.20

CC-1 112 55.6 0.2 500 2.33 3.07

CC-2a 117 55.8 0.33 817 1.96 6.64

CC-2b 170 56.5 0.33 817 9.13 7.39

CC-3 168 56.2 0.45 1114 3.57 1.68

Test procedure and set up

The fire tests were performed using a gas furnace in the Structures Laboratory at the

Western Sydney University, Australia. Prior to fire exposure, the concrete specimen

was installed in the furnace as shown in Figure 7.2. For each specimen, two LVDTs

at 180º apart at the top and two at the bottom were used to measure the axial

shortening during the tests. The exact locations of the LVDTs are shown in Figure

7.2. For the fire resistance test, the procedure proposed by Tao and Ghannam (Tao,

Ghannam et al. 2016) was adopted as follows: After column was fixed in position,

two cameras were mounted onto the two windows of the furnace’s door. The camera

zoom was adjusted so that the specimen was in clear view of the cameras. Additional

Page 175: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

160

light source close to each window was provided to help the camera to capture the

images during the fire test. Thermocouples were connected to the thermocouple data

readers and the furnace door was locked firmly prior to the fire test. The test was

conducted through the following steps.

Figure 7.2 Fire test set-up

(1) Ambient testing phase. A predetermined axial compression load (N0) was

applied to the specimen. The furnace chamber had a floor area of 640 mm in

width and 630 mm in depth, and a height of 880 mm. Since the installed

specimens were higher than the furnace chamber, the height of the specimen

subjected to fire was 880mm. Pin-to-pin end boundary conditions were

applied to the column specimen using two cylindrical hinges and a loading

jack with a capacity of 2,000 kN, located at the top of furnace.

(2) Temperature rise phase. The applied load was kept constant during this phase

and the temperature was increased at an average rate of 40℃/min. The

Page 176: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

161

temperature inside the furnace was controlled by four gas burners. Five

thermocouples were used to monitor and record the furnace temperature.

(3) Constant temperature phase. After the furnace temperature reached 800 ℃,

the furnace temperature was maintained constant until the column failed to

support the applied axial load, and the column reached its ultimate state. The

failure criterion for axially loaded elements specified in ISO 834 standard

(1999) was adopted to identify the ultimate state of the column specimen in

the fire. ISO 834 standard failure criterion is either the axial shortening

exceeding 0.01L mm or the deformation rate exceeding 0.003L mm/min,

where L is the length of the specimen in mm. The criterion for elapsed time

from the beginning of the temperature rise phase to the time of failure were

satisfied. This was defined as the fire resistance (tr) of the column specimen.

It should be noted that the ISO 834-1 (1999) fire curve was not used due to the

limitations of the furnace temperature, which could only reach a specified maximum

temperature of 800°C.

Test Results and Discussion

General observations and failure modes

The images captured by the two cameras that were mounted onto the two windows of

the furnace’s door, at different times during the fire tests are presented in Figures

7.3-7.12. The use of different types of concrete had obvious influence on the spalling

behaviour, while no spalling was observed for OGC columns, explosive cover

spalling was observed in all of HGC columns and one of CC columns (CC-2a). The

first spalling was observed at around 34 min for HGC-1a, 43 min for HGC-1b and 25

Page 177: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

162

min for HGC-2 after the heating was started. In the mentioned times, the surface

temperature was around 680 ℃ for HGC-1a, 710 ℃ for HGC-1b and 716 ℃ for

HGC-2. Also, for CC-2a column, the first spalling was observed at around 26 min

after the heating was started while the surface temperature was 655℃. In these

columns, the spalling was further extended as seen in Figures 6-8 and Figure 10 and

resulted in earlier failure of these columns.

The general failure of the OGC, HGC and CC columns are depicted in Figures 7.13-

7.15, respectively and their different views are shown in Figures 7.16-7.18,

respectively. In general, all columns failed due to global buckling accompanied by

local buckling concentrated around the mid-height of each column. For OGC

columns as seen in Figures 7.13 and 7.16, the severity of the damage was similar,

while some cracks initiated at mid-height of the columns and progressed and

widened with time and caused the failure with no cover spalling during the test. For

HGC columns as shown in Figures 7.14 and 7.17, explosive spalling observed which

exposed the reinforcement to fire and accordingly caused earlier failure of the

columns. The spalling of HGC could be due to their dense microstructure which

prohibits the release of vapour and increases the pore pressure. For CC columns as

seen in Figures 7.15 and 7.18, the spalling behaviour was affected by the initial load

level and age of concrete. With the same load level for CC-2a and CC-2b, explosive

spalling occurred for CC-2a column at around 26 min from the start of heating while

no spalling was observed for CC-2b which could be due to the difference of age and

less moisture content for this column.

Page 178: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

163

Figure 7.3 OGC-1a column during the fire test at different times: (a) beginning of the test;

(b) 144 min after the start of fire exposure and (c) end of the test.

Figure 7.4 OGC-1b column during the fire test at different times: (a) beginning of the test;

(b) 165 min after the start of fire exposure and (c) end of the test.

Page 179: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

164

Figure 7.5 OGC-2 column during the fire test at different times: (a) beginning of the test; (b)

78 min after the start of fire exposure and (c) end of the test.

Figure 7.6 HGC-1a column during the fire test at different times: (a) beginning of the test;

(b) 34 min after the start of fire exposure and (c) end of the test.

Page 180: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

165

Figure 7.7 HGC-1b column during the fire test at different times: (a) beginning of the test;

(b) 43 min after the start of fire exposure; (c) 48 min after the start of fire exposure and (d)

end of the test.

Figure 7.8 HGC-2 column during the fire test at different times: (a) beginning of the test; (b)

25 min after the start of fire exposure; (c) 32 min after the start of fire exposure and (d) end

of the test.

Page 181: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

166

Figure 7.9 CC-1 column during the fire test at different times: (a) beginning of the test; (b)

45 min after the start of fire exposure; (c) 265 min after the start of fire exposure and (d) end

of the test.

Figure 7.10 CC-2a column during the fire test at different times: (a) beginning of the test;

(b) 26 min after the start of fire exposure and (c) 42 min after the start of fire.

Page 182: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

167

Figure 7.11 CC-2b column during the fire test at different times: (a) beginning of the test;

(b) 46 min after the start of fire exposure and (c) end of the test.

Figure 7.12 CC-3 column during the fire test at different times: (a) beginning of the test; (b)

43 min after the start of fire exposure and (c) end of the test.

Page 183: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

168

Figure 7.13 A general view of the OGC columns after test

Figure 7.14 A general view of the HGC columns after test

Page 184: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

169

Figure 7.15 A general view of the CC columns after test

Figure 7.16 Different views of OGC columns after failure

Page 185: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

170

Figure 7.17 Different views of HGC columns after failure

Figure 7.18 Different views of CC columns after failure

Page 186: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

171

Temperature versus time curves

The measured temperature (T) versus time (t) curves for OGC, HGC and CC

columns are illustrated in Figures 7.19-7.21, respectively. The average furnace

temperature is also shown in the figures for each tested column. It should be

mentioned that some thermocouples failed during the fire test and thus the

corresponding measured temperatures are not presented.

In all samples, the temperature inside the columns increased rapidly to around 100 ℃

and beyond this temperature, the rate of increase was slower. This behaviour is

associated with the moisture migration towards the centre of column due to exposure

to high temperature (Kodur, Cheng et al. 2003).

In general, different samples of the same concrete type showed almost similar

behaviour in terms of temperature distribution at various points. However, different

patterns were observed between columns with different concrete types. For all

specimens, point 1 located at the surface of the columns showed higher temperature

than the others followed by point 4 which was the steel reinforcement temperature.

The temperature distribution at various section depths was significantly influenced

by the thermal properties of concrete mixes and thus OGC, HGC and CC columns

showed different temperature distribution to each other.

The temperature distribution against time at different points for OGC-2, HGC-1a and

CC-2b columns are compared in Figure 7.22. As it can be seen, in general, the

temperature rise in OGC-2 at all points was slower than the others due to its lower

thermal conductivity as observed in former thermal properties tests. For instance, at a

heating time of around 46 minutes at corresponding furnace temperature of 800 ℃

Page 187: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

172

for all samples, the measured temperatures of points 2, 3 and 4 were 130, 74 and 180

℃ for OGC-2, 351, 118 and 477 ℃ for HGC-1a and 190, 122 and 360 ℃ for CC-2b.

It seems that the heat transfer in OGC-2 was the slowest among these columns. It

should be mentioned that although HGC showed lower thermal conductivity

compared to CC, the measured temperatures were higher which was due to the

severe spalling that occurred in this column and as a result exhibited the fastest heat

transfer.

Figure 7.19 Temperature T versus time t curves of OGC columns

Page 188: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

173

Figure 7.20 Temperature T versus time t curves of HGC columns

Figure 7.21 Temperature T versus time t curves of CC columns

Page 189: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

174

Figure 7.22 Effect of concrete type on the temperature distribution of column section at

different points: (a) point 2; (b) point 3; (c) point 4.

Axial deformation

The axial deformation (∆) versus time (t) curves for all tested columns are presented

in Figure 7.23. All the columns showed similar ∆-t behaviour divided into three

stages. All the curves showed an initial expansion due to the thermal expansion of

concrete and steel, followed by a gradual contraction after the expansion peak due to

the yielding of the reinforcement. Finally, with time the axial deformation increased

significantly and the columns failed and could not support the applied axial load.

Figure 7.24 depicts the effect of concrete type on the axial behaviour of columns

subjected to fir at the same load. As seen in Figures 7.23 and 7.24, concrete type and

load level had remarkable influence on the shape of ∆-t curve at different stages and

fire resistance of the columns. As expected, with the same concrete type, the columns

Page 190: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

175

subjected to lower initial load level showed longer fire resistance. The axial

deformation in the expansion stage was very different for columns with different

concrete type. The OGC columns maintained the expansion stage for a considerable

duration before contracting as well as deformed larger under expansion compared to

those of CC and HGC columns due to their lower thermal conductivity. The thermal

expansion period (min) and its corresponding deformation (mm) for different

columns were as follows: OGC-1: 71, 2.09; OGC-2: 53, 1.3; OGC-3: 150, 2.1; HGC-

1: 35, 0.8; HGC-2: 30, 0.5; HGC-3: 36, 1.1 , ; CC-1: 51, 1.2; CC-2: 28, 0.2; CC-3:

21, 0.4; CC-4: 21, 0.6. In general, with the same heating rate and load level, OGC

columns exhibited larger axial deformation with time compared to the other columns

which shows that the use of OGC increases the ductility of reinforced columns

subjected to fire. The axial deformations for the columns were as follows: OGC-1:

10.87, OGC-2: 12.16, OGC-3: 5.73, HGC-1: 0.74, HGC-2: 0.75, HGC-3: 0.80, CC-

1: 15.58, CC-2: 4.35, CC-3: 8.8, CC-4: 7.8 mm. Detailed analysis will be presented

in the following subsection.

Page 191: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

176

Figure 7.23 Axial deformation versus time curves for OGC, HGC and CC columns

Figure 7.24 Comparison of axial deformation versus time curves for the columns with

similar initial load levels: (a) Initial load level of 0.33; (b) Initial load level of 0.45.

Fire resistance

The measured fire resistance (tR) for each column specimen is presented is Table 7.2.

It should be mentioned that for columns CC-1 (load level 0.20) and OGC-3 (load

level 0.3) the heating rate was different from the other specimens due to some

Page 192: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

177

problem to control the temperature of the furnace. As expected, the fire resistance of

the RC columns was significantly affected by the load level and concrete type. With

increasing load level, the fire resistance of the columns decreased. With similar load

level, OGC columns exhibited significantly higher fire resistance compared to HGC

and CC columns. It should be mentioned that the occurrence of spalling in CC-2a,

decreased the fire resistance by 28 min compared to CC-2b with the same load level.

Also HGC columns showed very low fire resistance due to the explosive spalling

which caused their earlier failure. On average, the fire resistance of OGC columns

was around 40% higher than CC columns with load level of 0.33 and 28% higher

than CC columns with load level of 0.45. Apparently, OGC reinforced columns had

the best fire resistance among columns with three types of concrete. As mentioned

before, this superior performance could be due to the lower thermal conductivity of

OGC which in turn resulted in delay of heat development in the columns. A finding

was also reached in Chapter 5 that cylinders made with OGC also achieved better

fire resistance than cylinders made with HGC and CC. However, further research is

required to address the reason for spalling and poor fire performance of hybrid OPC-

geopolymer concrete as previous studies mostly focused on the fire performance of

conventional geopolymer mixes.

Page 193: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

178

Table 7.2 Measured Fire resistance of the columns

Specimen

label

Load level

nf

tR

(min)

OGC-1a 0.33 186.20

OGC-1b 0.33 198.50

OGC-2 0.45 150.25

HGC-1a 0.33 56.05

HGC-1b 0.33 50.53

HGC-2 0.45 38.10

CC-1 0.20 271.58

CC-2a 0.33 87.78

CC-2b 0.33 115.98

CC-3 0.45 108.36

Evaluation of Test Results with FE Predictions

ABAQUS finite element (FE) software was utilised to build a model to simulate the

behaviour of the tested columns. The sequentially coupled thermal stress analysis

available in ABAQUS was used to simulate the fire test as also recommended in

previous studies (Han, Chen et al. 2013, Tao, Ghannam et al. 2016). In this method,

the stress analysis is dependent on the thermal analysis but the thermal analysis is

independent of the stress analysis. The temperature field of the column is assigned to

the stress analysis through a predefined field. In this method, the elements in the

thermal analysis and stress analysis must be of the same family type, and it is

preferable to use the same mesh size in both models to overcome convergence

problems. The details of thermal and stress analysis is presented in this chapter.

Page 194: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

179

Thermal analysis

The measured thermal properties presented in Chapter 5 were used to predict the

temperature distribution in the columns. The heat transfer analysis was used to

predict heat flow and temperature in the specimens.

7.4.1.1 Material properties

For thermal analysis, material properties of concrete including thermal conductivity,

density and specific heat were defined according to conducted experiments reported

in Chapter 5. Thermal properties of reinforcing steel at elevated temperatures were

defined according to Eurocode 4 (2005) including thermal conductivity and specific

heat as shown in Figures 7.25 and 7.26, respectively.

Figure 7.25 Thermal conductivity of steel as a function of temperature according to

Eurocode 4 (2005)

0

10

20

30

40

50

60

0 200 400 600 800 1000 1200 1400

Ther

mal

co

nd

uct

ivit

y (W

/mk)

Temperature (ºC )

Page 195: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

180

Figure 7.26 Specific heat of steel as a function of temperature according to Eurocode 4

(2005)

7.4.1.2 Analysis type

Heat Transfer analysis was used to analyse heat and temperature distribution. The

temperature versus time was the main input of the analysis which was the same as

the recorded furnace temperature during the test. Analysis increments and final

temperature were set accordingly.

7.4.1.3 Boundary conditions and interaction

The interaction between concrete and steel reinforcements was defined by tie

constraint. Surface radiation and surface film condition were defined for heated and

isolated part of the column. In addition, initial room temperature was defined by

defining a predefined field value for the whole column.

Stress analysis

The stress analysis was divided into two consecutive steps. In the first step, the axial

load was applied to the column at ambient temperature. The second step included the

0

500

1000

1500

2000

2500

3000

0 200 400 600 800 1000 1200 1400

Spec

ific

hea

t (J

/kgK

)

Temperature (ºC)

Page 196: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

181

loading from the first step and the temperature from the thermal analysis which has

to be introduced to the stress analysis as a predefined field through the second step.

Geometric nonlinearity was taken into account in the stress analysis.

7.4.2.1 Material properties

The thermal expansion of concrete at elevated temperatures was calculated according

to the model proposed by Eurocode 2 (2005) for siliceous aggregate as given by

Equations 7.1:

For siliceous aggregate:

for 20℃ ≤ 𝑇 ≤ 700℃ ∆𝑙

𝑙= −1.8 × 10−4 + 9 × 10−6𝑇 + 2.3 × 10−11𝑇3

for 700℃ < 𝑇 ≤ 1200℃ ∆𝑙

𝑙= 14 × 10−3 (7.1)

where T is the concrete temperature (℃).

In terms of compressive strength at elevated temperatures, the values measured in

Chapter 5 were used. Concrete damage plasticity is used in ABAQUS to define the

tensile and compressive behaviour of concrete. In this model, the plasticity, concrete

compressive and tensile behaviours should be defined. For plasticity, all default

values of parameters proposed by ABAQUS were adopted, with the exception of the

dilation angle as follows: The ratio of the second stress invariant on the tensile

meridian to that on the compressive meridian (Kc) was set to 2/3; flow potential

eccentricity (e) equal to 0.1; and the ratio of the compressive strength under biaxial

loading to uniaxial compressive strength (fb0 / fc) equal to 1.16.

The dilation angle (ψ) was set at 40° in this model based on the sensitivity analysis

carried out in this thesis. The “GFI” option was chosen to define the tensile

Page 197: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

182

behaviour by entering the tensile stress and fracture energy. Fracture energy at

ambient temperature (GF) was calculated using Equation 7.2, proposed by Gao et al.

(Gao, Dai et al. 2013) which showed that fracture energy is not dependent on

temperature as follows:

𝐺𝐹 = (0.0469𝑑𝑚𝑎𝑥2 − 0.5𝑑𝑚𝑎𝑥 + 26)(

𝑓𝑐′

10)0.7 (7.2)

where 𝑑𝑚𝑎𝑥 is the maximum aggregate size (mm) and 𝑓𝑐′ is the concrete cylinder

strength (MPa).

The tensile strength of concrete at high temperature was calculated by considering

the high-temperature reduction factors proposed by Eurocode 2 as depicted in Figure

7.27.

Figure 7.27 Concrete tensile reduction factor versus temperature Eurocode 2 (2005)

For carbon steel, Eurocode 3 (2005) was used to model the stress-strain and thermal

expansion as presented in Table 7.3.

Page 198: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

183

Table 7.3 Eurocode 3 parameters and formulations for carbon steel properties

Strain range Stress f Tangent modulus Et

𝜀 ≤ 𝜀𝑝,𝑇 𝜀. 𝐸𝑠,𝑇 𝐸𝑠,𝑇

𝜀𝑝,𝑇 ≤ 𝜀 ≤ 𝜀𝑦,𝑇 𝑓𝑝,𝑇 − 𝑐 + (

𝑏

𝑎)√𝑎2 − (𝜀𝑦,𝑇 − 𝜀)2

𝑏(𝜀𝑦,𝑇 − 𝜀)

𝑎 × √𝑎2 − (𝜀𝑦,𝑇 − 𝜀)2

𝜀𝑦,𝑇 ≤ 𝜀 ≤ 𝜀𝑡,𝑇 𝑓𝑦,𝑇 0

𝜀𝑡,𝑇 ≤ 𝜀 ≤ 𝜀𝑢,𝑇 𝑓𝑦,𝑇(1 −

(𝜀 − 𝜀𝑡,𝑇)

(𝜀𝑢,𝑇 − 𝜀𝑡,𝑇))

-

𝜀 = 𝜀𝑢,𝑇 0 -

parameters 𝜀𝑝,𝑇 = 𝑓𝑝,𝑇/𝐸𝑠,𝑇 𝜀𝑦,𝑇 = 0.02 𝜀𝑡,𝑇 = 0.15 𝜀𝑢,𝑇 = 0.2

Functions

𝑎2 = (𝜀𝑦,𝑇 − 𝜀𝑝,𝑇)(𝜀𝑦,𝑇 − 𝜀𝑝,𝑇 +𝑐

𝐸𝑠,𝑇)

𝑏2 = 𝑐(𝜀𝑦,𝑇 − 𝜀𝑝,𝑇)𝐸𝑠,𝑇 + 𝑐2

𝑐 =(𝑓𝑦,𝑇 − 𝑓𝑝,𝑇)2

(𝜀𝑦,𝑇 − 𝜀𝑝,𝑇)𝐸𝑠,𝑇 − 2(𝑓𝑦,𝑇 − 𝑓𝑝,𝑇)

where f y,T : effective yield strength, f p,T : proportional limit, E s,T : slope of the linear

elastic range, ε p,T : strain at the proportional limit, ε y,T : yield strain, ε u,T : ultimate

strain.

For temperature below 400℃, an alternative strain-hardening formula could be used

as follows;

For 0.02 < 𝜀 < 0.04 𝑓𝑠 = 50(𝑓𝑢,𝑇 − 𝑓𝑦,𝑇)𝜀 + 2𝑓𝑦,𝑇 − 𝑓𝑢,𝑇

For 0.04 < 𝜀 < 0.15 𝑓𝑠 = 𝑓𝑢,𝑇

Page 199: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

184

For 0.15 < 𝜀 < 0.20 𝑓𝑠 = 𝑓𝑢,𝑇(1 − 20(𝜀 − 0.15))

For 𝜀 ≥ 0.2 𝑓𝑠 = 0 (7.3)

where f u,T is the ultimate strength at elevated temperatures, allowing for strain-

hardening.

The ultimate strength at elevated temperatures, allowing for strain-hardening, should

be determined as given by Equation 7.4:

for Ts < 300 °C f u,T = 1,25 f y,T

for 300 °C ≤ Ts < 400 °C f u,T = f y,T (2 - 0.0025 Ts )

for Ts ≥ 400 °C f u,T = f y,T (7.4)

The relative thermal elongation of steel (Δl/l) was determined according to Eurocode

3 as indicated below in Equation 7.5:

for 20 °C ≤ Ts < 750 °C Δl / l = 1.2 × 10-5 Ts + 0.4 × 10-8 Ts2 – 2.416 × 10-4

for 750 °C ≤ Ts ≤ 860 °C Δl / l = 1.1 × 10-2

for 860 °C < Ts ≤ 1200 °C Δl / l = 2 × 10-5 Ts – 6.2 × 10-3 (7.5)

where l is the length at 20 °C; Δl is the temperature-induced elongation; and Ts is the

steel temperature in °C.

7.4.2.2 Element divisions and boundary conditions

As observed form Figure 7.2, the tested columns were pin-ended. To account for the

difference in thermal expansion of carbon steel and concrete, the upper boundary

condition was modelled by defining a rigid plate at the upper end of the column. The

Page 200: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

185

plate is controlled by a reference point located at its centroid. A pinned-boundary

condition was assigned to this point. Contact interaction was defined between the

plate and the top of the concrete. At the lower end of the column, a pinned-boundary

condition was assigned to a reference point located at the centroid of the lower end.

This reference point was coupled with the lower end of the column so that they had

the same deformation.

For columns with reinforcing bars, tie constraint was assumed between the

reinforcing bars and the concrete core. Details of the model are shown in Figure 7.28.

The whole column section and length were used in the model.

Figure 7.28 Finite element model for stress analysis

Comparison of predictions with experimental results

In order to validate the accuracy of the established FE models, the models were used

to simulate the test data reported in this thesis. For the fire-resistance tests, the failure

modes, temperature versus time curves and axial deformation versus time curves

Page 201: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

186

were predicted by the FE models. It should be mentioned that spalling of concrete

was not considered in FE model and further research is required in future to take

account of this model.

The predicted versus measured temperature distribution as a function of time curves

of the reinforced columns at different points are depicted in Figures 7.29 -7.31. In

general, there is a reasonably good agreement between measured and calculated

temperatures at different points for OGC and CC columns. However, some

differences between the two sets of patterns might be due to variability of

thermocouple positions at the time of concrete casting. However, for HGC columns,

as spalling was not considered in the modelling, considerable difference between

measured and predicted temperature distribution at different points was observed.

This behaviour was due to the fact that the occurrence of spalling exposed the inner

layers to higher temperatures compared to FE model and thus higher temperature

was measured.

The other reason for the difference between the measured and predicted temperature

especially at the centre of columns at early stages of fire exposure, is attributed to the

thermally-induced migration of moisture towards the centre of the columns which

was not taken into account in FE modelling (Kodur, Wang et al. 2004). Thus, the

temperature was increased rapidly for tested graphs followed by a period of almost

constant temperatures at about 100 ℃ which was not seen in predicted graphs.

Further research is required to take into account the evaporation of moisture and

migration of moisture toward the centre in numerical modelling.

Page 202: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

187

Figure 7.29 Tested versus measured Temperature as a function of Time for OGC columns

Figure 7.30 Tested versus measured Temperature as a function of Time for HGC columns

Page 203: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

188

Figure 7.31 Tested versus measured Temperature as a function of Time for CC columns

The comparison between measured and predicted axial displacement (∆) versus time

(t) curves of the reinforced columns subjected to elevated temperatures are shown in

Figures 7.32 -7.34. It can be seen that the overall axial behaviour of columns with

time was predicted accurately with FE analysis. However, the variation between the

measured and predicted deformations could be somewhat due to the spalling

especially for HGC columns. As reported previously, the extent of spalling has a

significant effect on axial deformations of reinforced columns exposed to fire and

less difference between the measured and predicted axial deformation graphs is

observed with lower percentage of spalling (Kodur, Wang et al. 2004).

The difference between measured and predicted axial deformation at the expansion

stage was 1.7, 1.1 and 1.0 for OGC-1a, OGC-1b and OGC-2, 0.8 , 1.0 and 0.43 for

HGC-1a, HGC-1b and HGC-2, 0.8, 0.3, 0.4 and 0.5 for CC-1, CC-2a, CC-2b and

Page 204: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

189

CC-3, respectively. The differences are considered relatively small with respect to

the length of columns.

In the descending branch of the axial deformation curve of some columns such as

HGC columns, the predicted values were larger than the measured deformations

which might be due to the occurrence of spalling in these columns. It should be noted

that the axial deformation is also influenced by other different factors including load,

thermal expansion, bending and the effect of transit creep, etc and thus these

parameters should be taken into account in modelling to achieve better agreement

(Kodur, Wang et al. 2004).

Figure 7.32 Comparison of axial deformation for OGC columns

Page 205: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

190

Figure 7.33 Comparison of axial deformation for HGC columns

Figure 7.34 Comparison of axial deformation for CC columns

Page 206: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

191

The measured and predicted fire resistance of columns are compared in Table 7.4.

The predicted fire resistance was within 12% for OGC columns, 48% for HGC

columns and 15% for CC columns, of the measured values. The large variation

between the measured and predicted fire resistance of HGC columns is due to the

neglect of spalling. The accuracy of the predictions could be improved in future, by

considering the influence of spalling in the model.

Table 7.4 Comparison of fire resistance of columns

Specimen

label

Load level

nf

Measured tR

(min)

Predicted tR

(min)

OGC-1a 0.33 186.20 209.17

OGC-1b 0.33 198.50 207.35

OGC-2 0.45 150.25 149.22

HGC-1a 0.33 56.05 83.10

HGC-1b 0.33 50.53 73.12

HGC-2 0.45 38.10 53.56

CC-1 0.20 271.58 240.46

CC-2a 0.33 87.78 96.30

CC-2b 0.33 115.98 110.70

CC-3 0.45 108.36 106.35

Conclusions

The fire performance of reinforced concrete columns with three different types of

concrete including OGC, HGC and CC was experimentally and numerically

Page 207: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

192

investigated in this chapter. To conclude this chapter, the important findings are

presented as follows:

• Concrete type had significant effect on the spalling behaviour of reinforced

columns. No spalling was observed for OGC columns, but explosive cover

spalling was observed in all HGC columns and one of CC columns.

• The temperature distribution with time at different section positions was

almost similar for columns with the same type of concrete. However,

columns with different types of concrete exhibited distinct temperature

distribution progression with time. In general, lower values of temperature

were recorded for OGC columns compared to those of HGC and CC columns

due to their lower thermal conductivity. Thus, the temperature transferred

faster in HGC and CC columns than OGC columns.

• The axial deformation of reinforced concrete columns exposed to fire was

found to be affected by the type of concrete and initial load level. On average,

OGC columns showed larger elongation values at expansion stage compared

to those of HGC and CC columns. Also, with similar heating rate and load

level, OGC columns exhibited larger axial deformation with time compared

to the other columns which shows that the use of OGC increased the ductility

of reinforced columns subjected to fire.

• The fire resistance of RC columns was substantially influenced by the initial

load level and concrete type. By increasing load level, the fire resistance of

the columns decreased. At a similar load level, on average, OGC columns

exhibited significantly higher fire resistance compared to HGC and CC

Page 208: Mechanical, Thermal and Structural Investigations of One

Chapter 7 Performance of Reinforced Concrete Columns at Elevated

Temperatures

193

columns. The fire resistance of OGC columns, on average, was around 40%

higher than CC columns with load level of 0.33 and 28% higher than CC

columns with load level of 0.45.

• ABAQUS finite element (FE) software was used to develop a model to

simulate the behaviour of the tested columns. In general, there is a reasonably

good agreement between measured and calculated temperatures at different

points for OGC and CC columns. However, for HGC columns, as spalling

was not considered in the modelling, considerable difference between

measured and predicted temperature distribution at different points was

observed. The overall axial behaviour of columns with time was predicted

accurately with FE analysis. However, the variation between the measured

and predicted deformations could be somewhat due to the spalling, especially

for HGC columns. Thus, more research is still required to improve the

accuracy of FE analysis by taking into account the spalling model and

moisture migration towards the centre of columns.

The use of one-part geopolymer concrete has been found to significantly improve the

fire performance of reinforced columns. It is necessary to carry out experiments in

future subjected to ISO 834 standard fire curve to clarify and determine the influence

of this type of concrete on the fire performance of reinforced columns and promote

its application in building industry.

Page 209: Mechanical, Thermal and Structural Investigations of One

Chapter 8 Conclusions and Future Research Needs

194

Chapter 8 Conclusions and Future Research Needs

Conclusions

This thesis presents a study on the behaviour of one-part geopolymer concrete and

hybrid OPC-geopolymer concrete at both ambient and elevated temperatures and

their application in reinforced concrete columns. Based on the experimental and

analytical investigations reported on this topic, conclusion drawn from the major

outcomes of this thesis are summarised as follows:

(1) The inclusion of OPC in geopolymer concrete mixes decreases the workability

and setting time as it accelerates the geopolymerisation reaction. To avoid the rapid

setting of concrete mixes with OPC content of 30% and above, a suitable amount of

citric acid could be used to retard the OPC hydration.

(2) The one-part hybrid OPC-geopolymer concrete mixes with solid potassium

carbonate as activator achieved higher strength compared to those of the control

mixes without activator especially at early ages. By inclusion of 10% or higher

content of OPC in the binder of geopolymer mixes, the 28-day compressive strength

of higher than 30 MPa is achievable which ensure the feasibility of using this type of

concrete in structural applications. Microstructural investigation showed that hybrid

OPC-geopolymer mixes produces a mixture of cross-linked network of geopolymeric

and CSH gels in the binder system, which allows the concrete to develop high early

and ultimate strength.

(3) The mechanical and microstructural properties of one-part geopolymers are

significantly affected by type/combination of aluminosilicate precursors and solid

Page 210: Mechanical, Thermal and Structural Investigations of One

Chapter 8 Conclusions and Future Research Needs

195

activators. The inclusion of slag in one-part fly ash mixes results in more compact

and denser structure and accordingly higher early and final strength is achievable

without the need for heat curing.

(4) A relatively high compressive strength of 38 MPa is obtained when sodium

silicate, calcium hydroxide and lithium hydroxide are combined and used as activator

in fly ash/slag based geopolymer.

(5) With similar content of aggregate and binder and also similar water to binder

ratio, one-part geopolymer and hybrid OPC-geopolymer concrete mixes achieved the

28-day compressive strength of 40.4 and 42.2 MPa, respectively compared to 52.5

MPa for OPC concrete. Also, lower values in terms of modulus of elasticity,

modulus of rapture and indirect tensile strength for these mixes compared to OPC

concrete were obtained.

(6) At elevated temperatures, one-part geopolymer and hybrid OPC-geopolymer

concrete mixes exhibit different behaviour in comparison with OPC concrete. At 800

°C, one-part geopolymer and hybrid OPC-geopolymer concrete displayed higher

strength of about 44% and 38% of their initial strength, respectively compared to

26% for OPC samples. Furthermore, the strain at peak strength at 800 °C was much

higher for one-part geopolymer concrete compared to those of hybrid OPC-

geopolymer and OPC concrete.

(7) The thermal properties of concrete at ambient and elevated temperatures are

substantially influenced by the type of concrete. The thermal conductivity of hybrid

OPC-geopolymer and OPC concrete samples decreased up to 400 °C and then

remained almost stable while for one-part geopolymer concrete, the thermal

conductivity slightly decreased between 23 and 200 °C and was constant beyond this

Page 211: Mechanical, Thermal and Structural Investigations of One

Chapter 8 Conclusions and Future Research Needs

196

range. In general, at all temperatures, one-part geopolymer concrete had the lowest

thermal conductivity compared to other concrete mixes.

(8) Comparison of the measured modulus of elasticity and stress-strain curves for

one-part geopolymer and hybrid OPC-geopolymer concrete with the existing models

showed that these predictions give very good predictions for these types of concrete.

(9) The use of Eurocode and ASCE models for predicting the compressive strength

of one-part geopolymer and hybrid OPC-geopolymer concrete is very conservative.

Furthermore, the Eurocode predictions cannot be used to predict the thermal

conductivity of one-part geopolymer concrete at elevated temperatures.

(10) The behaviour of axially loaded reinforced concrete columns is affected by the

compressive strength of concrete and load eccentricity. The use of hybrid OPC-

geopolymer concrete with higher compressive strength in reinforced column,

resulted in a substantial increase of ultimate capacity compared to that of reinforced

one-part geopolymer concrete column. The variation of concrete strength was found

to have more influence on the ultimate load capacity of reinforced concrete columns

than load eccentricity.

(11) The developed FE model in ABAQUS, can predict the test results of reinforced

columns with different types of concrete such as one-part geopolymer and hybrid

OPC-geopolymer concrete reasonably well.

(12) The type of concrete has significant effect on the fire performance of reinforced

concrete columns. No spalling was occurred for any of the one-part geopolymer

concrete columns while substantial spalling was observed for all of the hybrid OPC-

geopolymer concrete columns and one of the OPC concrete columns.

Page 212: Mechanical, Thermal and Structural Investigations of One

Chapter 8 Conclusions and Future Research Needs

197

(13) The temperature distribution with time at different cross section depths of

columns was different for three mixes of one-part geopolymer, hybrid OPC-

geopolymer and OPC concrete. Generally, at all temperatures, lower values were

recorded for one-part geopolymer concrete columns compared to those of hybrid

OPC-geopolymer and OPC concrete columns due to their lower thermal

conductivity.

(14) The reinforced one-part geopolymer concrete columns exhibited larger axial

deformation with time compared to the other columns with similar load ratio, which

shows that the use of one-part geopolymer concrete increased the ductility of

reinforced concrete columns subjected to fire.

(15) The fire resistance of reinforced one-part geopolymer concrete columns, on

average, was around 40% higher than OPC concrete columns with load level of 0.33

and 28% higher than OPC concrete columns with load level of 0.45.

(16) The developed FE model in this study, can predict the test results of reinforced

columns with different types of concrete such as one-part geopolymer and hybrid

OPC-geopolymer concrete reasonably well.

(17) The developed FE model in this study provides reasonable prediction of fire

resistance of reinforced one-part geopolymer and OPC concrete columns.

Recommendations for Future Works

In spite of the extensive investigations being conducted on the behaviour of one-part

geopolymer concrete and hybrid OPC-geopolymer concrete and their application in

reinforced concrete columns at both ambient and elevated temperatures, further

research is necessary for a better understanding of one-part geopolymer concrete and

Page 213: Mechanical, Thermal and Structural Investigations of One

Chapter 8 Conclusions and Future Research Needs

198

hybrid OPC-geopolymer concrete, as well as more comprehensive application in

reinforced concrete columns. Thus, the following are suggestions for further research

needs:

(1) In this study the mechanical properties of one-part geopolymer concrete and

hybrid OPC-geopolymer concrete were addressed. However, further research is

required to investigate the long-term performance, durability and efflorescence of

these mixes. Also, the long term behaviour of reinforced concrete columns with one-

part geopolymer concrete and hybrid OPC-geopolymer concrete needs to be

investigated to clarify the durability of these types of concrete in reinforced columns.

(2) The developed one-part geopolymer concrete and hybrid OPC-geopolymer

concrete mixes did not achieve very high strength. Thus, more experimental work is

required to develop concrete mixes with higher strength by optimising the dosage

and combination of solid activators.

(3) The microstructural characteristics of one-part geopolymer concrete and hybrid

OPC-geopolymer concrete at ambient temperature were investigated. Further

research still needs to be conducted to evaluate the microscopic texture ad chemistry

alteration of these concrete mixes before and after fire exposure. Such knowledge is

essential to guide future efforts in improving concrete performance and to facilitate

the application of concrete in practice.

(4) The behaviour of one-part geopolymer concrete and hybrid OPC-geopolymer

concrete was mainly studied at ambient and elevated temperatures. However, the

post-fire behaviour of concrete is also worth investigating for a better understanding

of fire research on these types of concrete.

Page 214: Mechanical, Thermal and Structural Investigations of One

Chapter 8 Conclusions and Future Research Needs

199

(5) More research and experiments are required on the performance of reinforced

one-part geopolymer concrete and hybrid OPC-geopolymer concrete columns by

considering various parameters such as various load eccentricities and effect of

coupled compression and moment.

(6) The accuracy of FE analysis for prediction of the fire performance of reinforced

concrete columns needs to be improved by taking into account the spalling model

and moisture migration towards the centre of columns.

(7) Design of reinforced concrete columns with one-part geopolymer concrete and

hybrid OPC-geopolymer concrete should be studied in future to promote the

application of these types of concrete for engineering practice

Page 215: Mechanical, Thermal and Structural Investigations of One

References

200

References

(1992). "Structural fire protection. Manual No. 78." ASCE Committee on Fire

Protection, Structural Division, American Society of Civil Engineers, New York,

USA.".

(1993). "ACI Committee 363, State of the art of high strength concrete. American

Concrete Institute, Detroit, USA.".

(1995). "RILEM TC 129-MHT, Test methods for mechanical properties of concrete

at high temperatures—compressive strength for service and accident conditions."

Materials and Structures 28(3): 410-414.

(1997). "Standards Australia. Method of testing concrete: Determination of the Static

Chord Modulus of Elasticity and Poisson’s Ratio of Concrete Specimens.AS

1012.17; Sydney, Australia.".

(1998). "Standards Australia. Method of testing concrete: determination of properties

related to the consistency of concrete – slump test, AS 1012.3.1.".

(1999). "ASTM C 1437. Standard test methods for flow of hydraulic cement mortar.

Document Number: ASTM C1437-01, ASTM International.".

(1999). "ISO 834-1. Fire-resistance tests-Elements of building construction-Part 1 :

General requirements." International Organization for Standardization, Geneva.

(1999). "Standards Australia. Method of testing concrete: Determination of the

compressive strength of concrete specimen.AS 1012.9; Sydney, Australia.".

(2000). "Standards Australia. Methods of testing concrete: Determination of Indirect

Tensile Strength of Concrete Cylinders (’Brazil’ or Splitting Test).AS 1012.10;

Sydney, Australia.".

(2000). "Standards Australia. Methods of testing concrete: Determination of the

Modulus of Rupture.AS 1012.11; Sydney, Australia.".

(2005). "Design of concrete structures, Part 1-2: General rules-Structural fire design.

European Committee for Standardization, Brussels,Belgium.".

(2005). "Design of steel structures, Part 1–2: General rules-Structural fire design,

European Committee for Standardization, Brussels, Belgium.".

(2005). "Eurocode 4.Design of composite steel and concrete structures, Part1-2,

General rules of structural fire design.BSEN1994-1-2:2005.British Standards

Institution,London.".

(2007). "Standards Australia. Metallic materials - Tensile testing at ambinet

temperature.".

Page 216: Mechanical, Thermal and Structural Investigations of One

References

201

(2008). "ASTM C 191–08. Standard test methods for time of setting of hydraulic

cement by Vicat Needle. West Conshohocken, PA: ASTM International.".

(2009). "AS 3600: Concrete structures. Standards Australia,Sydney, NSW,

Australia.".

(2012). "Version 6.12." Dassault Systemes Simulia Corp, Rhode Island, USA.

Abdalqader, A. F., F. Jin and A. Al-Tabbaa (2016). "Development of greener alkali-

activated cement: utilisation of sodium carbonate for activating slag and fly ash

mixtures." Journal of Cleaner Production 113: 66-75.

Ahmad, S. H. and S. P. Shah (1985). "Structural properties of high strength concrete

and its implications for precast prestressed concrete." PCI Journal 30(6): 92-119.

Ahmed, S. and V. Vimonsatit (2012). "Compressive strength and thermal expansion

properties of geopolymer concrete at elevated temperature."

Ali, A. M., J. Sanjayan and M. Guerrieri (2017). "Performance of geopolymer high

strength concrete wall panels and cylinders when exposed to a hydrocarbon fire."

Construction and Building Materials 137: 195-207.

Aliabdo, A. A., A. E. M. A. Elmoaty and H. A. Salem (2016). "Effect of cement

addition, solution resting time and curing characteristics on fly ash based geopolymer

concrete performance." Construction and Building Materials 123: 581-593.

Ambily, P. S., K. Ravisankar, C. Umarani, J. K. Dattatreya and N. R. Iyer (2014).

"Development of ultra-high-performance geopolymer concrete." Magazine of

Concrete Research 66(2): 82-89.

Andalib, R., M. W. Hussin, M. Z. A. Majid, M. Azrin and H. H. Ismail (2014).

"Structural performance of sustainable waste palm oil fuel ash-fly ash geo-polymer

concrete beams." Journal of Environmental Treatment Techniques 2(3): 115-119.

Askarian, M., Z. Tao, G. Adam and B. Samali (2018). "Mechanical properties of

ambient cured one-part hybrid OPC-geopolymer concrete." Construction and

Building Materials 186: 330-337.

Assi, L., S. Ghahari, E. E. Deaver, D. Leaphart and P. Ziehl (2016). "Improvement of

the early and final compressive strength of fly ash-based geopolymer concrete at

ambient conditions." Construction and Building Materials 123: 806-813.

Atiş, C., E. Görür, O. Karahan, C. Bilim, S. Ilkentapar and E. Luga (2015). "Very

high strength (120 MPa) class F fly ash geopolymer mortar activated at different

NaOH amount, heat curing temperature and heat curing duration." Construction and

building materials 96: 673-678.

Bakharev, T. (2005). "Geopolymeric materials prepared using Class F fly ash and

elevated temperature curing." Cement and concrete research 35(6): 1224-1232.

Page 217: Mechanical, Thermal and Structural Investigations of One

References

202

Bakharev, T. (2006). "Thermal behaviour of geopolymers prepared using class F fly

ash and elevated temperature curing." Cement and Concrete Research 36(6): 1134-

1147.

Bell, J. L., P. E. Driemeyer and W. M. Kriven (2009). "Formation of ceramics from

metakaolin‐based geopolymers. Part II: K‐based geopolymer." Journal of the

American Ceramic Society 92(3): 607-615.

Bernal, S. A., J. L. Provis, R. J. Myers, R. San Nicolas and J. S. van Deventer (2015).

"Role of carbonates in the chemical evolution of sodium carbonate-activated slag

binders." Materials and Structures 48(3): 517-529.

Bratina, S., B. Čas, M. Saje and I. Planinc (2005). "Numerical modelling of

behaviour of reinforced concrete columns in fire and comparison with Eurocode 2."

International journal of solids and structures 42(21-22): 5715-5733.

Canfield, G. M., J. Eichler, K. Griffith and J. D. Hearn (2014). "The role of calcium in blended fly ash geopolymers." Journal of materials science 49(17): 5922-5933.

Cao, Y.-F., Z. Tao, Z. Pan and R. Wuhrer (2018). "Effect of calcium aluminate

cement on geopolymer concrete cured at ambient temperature." Construction and

Building Materials 191: 242-252.

Carmona, S. and A. Aguado (2012). "New model for the indirect determination of

the tensile stress–strain curve of concrete by means of the Brazilian test." Materials

and structures 45(10): 1473-1485.

Castel, A. and S. J. Foster (2015). "Bond strength between blended slag and Class F

fly ash geopolymer concrete with steel reinforcement." Cement and Concrete

Research 72: 48-53.

Cheah, C. B. and M. Ramli (2012). "Mechanical strength, durability and drying

shrinkage of structural mortar containing HCWA as partial replacement of cement."

Construction and Building Materials 30: 320-329.

Cheah, C. B. and M. Ramli (2013). "The engineering properties of high performance

concrete with HCWA–DSF supplementary binder." Construction and Building

Materials 40: 93-103.

Chen, C., Q. Li, L. Shen and J. Zhai (2012). "Feasibility of manufacturing

geopolymer bricks using circulating fluidized bed combustion bottom ash."

Environmental technology 33(11): 1313-1321.

Chindaprasirt, P., T. Chareerat, S. Hatanaka and T. Cao (2010). "High-strength

geopolymer using fine high-calcium fly ash." Journal of Materials in Civil

Engineering 23(3): 264-270.

Claeson, C. and K. Gylltoft (1998). "Slender high-strength concrete columns

subjected to eccentric loading." Journal of Structural Engineering 124(3): 233-240.

Page 218: Mechanical, Thermal and Structural Investigations of One

References

203

Collins, M. P. and D. Mitchell (1991). Prestressed concrete structures, Prentice Hall

Englewood Cliffs, NJ.

Collins, M. P., D. Mitchell and J. G. MacGregor (1993). "Structural design

considerations for high-strength concrete." Concrete international 15(5): 27-34.

Dattatreya, J., N. Rajamane, D. Sabitha, P. Ambily and M. Nataraja (2011). "Flexural

behaviour of reinforced Geopolymer concrete beams." International Journal of civil

and structural engineering 2(1): 138.

Dave, N. and V. Sahu (2013). "Approches towards industrial by-products utilization:

A review towards environmentally safe future." International Journal of Innovative

Research in Science, Engineering and Technology 2(7): 2847-2855.

Davidovits, J. (1989). "Geopolymers and geopolymeric materials." Journal of

Thermal Analysis and Calorimetry 35(2): 429-441.

Davidovits, J. (1989). "Geopolymers and geopolymeric materials." Journal of

thermal analysis 35(2): 429-441.

Davidovits, J. (1994). Properties of geopolymer cements. First international

conference on alkaline cements and concretes, Scientific Research Institute on

Binders and Materials Kiev State Technical University, Ukraine.

De Silva, P., K. Sagoe-Crenstil and V. Sirivivatnanon (2007). "Kinetics of

geopolymerization: role of Al2O3 and SiO2." Cement and Concrete Research 37(4):

512-518.

De Vargas, A. S., D. C. Dal Molin, A. C. Vilela, F. J. Da Silva, B. Pavao and H. Veit

(2011). "The effects of Na2O/SiO2 molar ratio, curing temperature and age on

compressive strength, morphology and microstructure of alkali-activated fly ash-

based geopolymers." Cement and concrete composites 33(6): 653-660.

Deb, P. S., P. Nath and P. K. Sarker (2014). "The effects of ground granulated blast-

furnace slag blending with fly ash and activator content on the workability and

strength properties of geopolymer concrete cured at ambient temperature." Materials

& Design (1980-2015) 62: 32-39.

Duxson, P., A. Fernández-Jiménez, J. L. Provis, G. C. Lukey, A. Palomo and J. Van

Deventer (2007). "Geopolymer technology: the current state of the art." Journal of

Materials Science 42(9): 2917-2933.

Duxson, P., A. Fernández-Jiménez, J. L. Provis, G. C. Lukey, A. Palomo and J. S.

van Deventer (2007). "Geopolymer technology: the current state of the art." Journal

of materials science 42(9): 2917-2933.

Duxson, P., G. C. Lukey and J. S. van Deventer (2007). "Physical evolution of Na-

geopolymer derived from metakaolin up to 1000 C." Journal of Materials Science

42(9): 3044-3054.

Page 219: Mechanical, Thermal and Structural Investigations of One

References

204

Duxson, P., J. L. Provis, G. C. Lukey, S. W. Mallicoat, W. M. Kriven and J. S. Van

Deventer (2005). "Understanding the relationship between geopolymer composition,

microstructure and mechanical properties." Colloids and Surfaces A:

Physicochemical and Engineering Aspects 269(1-3): 47-58.

Dwaikat, M. B. and V. Kodur (2009). "Hydrothermal model for predicting fire-

induced spalling in concrete structural systems." Fire Safety Journal 44(3): 425-434.

Feng, D., J. L. Provis and J. S. Deventer (2012). "Thermal Activation of Albite for

the Synthesis of One‐Part Mix Geopolymers." Journal of the American Ceramic

Society 95(2): 565-572.

Feng, X., M. Thomas, T. Bremner, B. Balcom and K. Folliard (2005). "Studies on

lithium salts to mitigate ASR-induced expansion in new concrete: a critical review."

Cement and Concrete Research 35(9): 1789-1796.

Fernández-Jiménez, A. and A. Palomo (2005). "Composition and microstructure of alkali activated fly ash binder: Effect of the activator." Cement and concrete research

35(10): 1984-1992.

Fernández-Jiménez, A., A. Palomo and M. Criado (2006). "Alkali activated fly ash

binders. A comparative study between sodium and potassium activators." Materiales

de Construcción 56(281): 51-65.

Franssen, J.-M. and J.-C. Dotreppe (2003). "Fire tests and calculation methods for

circular concrete columns." Fire technology 39(1): 89-97.

Ganesan, N., R. Abraham, S. Deepa Raj and K. Namitha (2015). "Effect of fibres on

the strength and behaviour of GPC columns." Magazine of Concrete Research 68(2):

99-106.

Ganesan, N., P. Indira and A. Santhakumar (2013). "Prediction of ultimate strength

of reinforced geopolymer concrete wall panels in one-way action." Construction and

building materials 48: 91-97.

Gao, W., J.-G. Dai, J. Teng and G. Chen (2013). "Finite element modeling of

reinforced concrete beams exposed to fire." Engineering structures 52: 488-501.

Gao, X., Q. Yu and H. Brouwers (2015). "Reaction kinetics, gel character and

strength of ambient temperature cured alkali activated slag–fly ash blends."

Construction and Building Materials 80: 105-115.

Garcia-Lodeiro, I., A. Palomo, A. Fernández-Jiménez and D. Macphee (2011).

"Compatibility studies between NASH and CASH gels. Study in the ternary diagram

Na2O–CaO–Al2O3–SiO2–H2O." Cement and Concrete Research 41(9): 923-931.

Görhan, G. and G. Kürklü (2014). "The influence of the NaOH solution on the

properties of the fly ash-based geopolymer mortar cured at different temperatures."

Composites part b: engineering 58: 371-377.

Page 220: Mechanical, Thermal and Structural Investigations of One

References

205

Greiser, S., G. Gluth, P. Sturm and C. Jäger (2018). "29Si {27Al}, 27Al {29Si} and

27Al {1H} double-resonance NMR spectroscopy study of cementitious sodium

aluminosilicate gels (geopolymers) and gel-zeolite composites."

Hajimohammadi, A., J. L. Provis and J. S. Van Deventer (2008). "One-part

geopolymer mixes from geothermal silica and sodium aluminate." Industrial &

Engineering Chemistry Research 47(23): 9396-9405.

Hajimohammadi, A., J. L. Provis and J. S. van Deventer (2010). "Effect of alumina

release rate on the mechanism of geopolymer gel formation." Chemistry of Materials

22(18): 5199-5208.

Hajimohammadi, A., J. L. Provis and J. S. van Deventer (2011). "The effect of silica

availability on the mechanism of geopolymerisation." Cement and Concrete

Research 41(3): 210-216.

Hajimohammadi, A. and J. S. van Deventer (2017). "Characterisation of one-part

geopolymer binders made from fly ash." Waste and Biomass Valorization 8(1): 225-

233.

Hajjaji, W., S. Andrejkovičová, C. Zanelli, M. Alshaaer, M. Dondi, J. Labrincha and

F. Rocha (2013). "Composition and technological properties of geopolymers based

on metakaolin and red mud." Materials & Design (1980-2015) 52: 648-654.

Han, L.-H., F. Chen, F.-Y. Liao, Z. Tao and B. Uy (2013). "Fire performance of

concrete filled stainless steel tubular columns." Engineering Structures 56: 165-181.

Hardjito, D. and B. V. Rangan (2005). "Development and properties of low-calcium

fly ash-based geopolymer concrete."

Hardjito, D., S. Wallah, M. Sumajouw and B. Rangan (2005). The stress-strain

behaviour of fly ash-based geopolymer concrete. Developments in Mechanics of

Structures and Materials, AA Balkema Publishers, The Netherlands.

Hardjito, D., S. E. Wallah, D. M. Sumajouw and B. V. Rangan (2004). "On the

development of fly ash-based geopolymer concrete." Materials Journal 101(6): 467-

472.

He, J., Y. Jie, J. Zhang, Y. Yu and G. Zhang (2013). "Synthesis and characterization

of red mud and rice husk ash-based geopolymer composites." Cement and Concrete

Composites 37: 108-118.

He, J., J. Zhang, Y. Yu and G. Zhang (2012). "The strength and microstructure of

two geopolymers derived from metakaolin and red mud-fly ash admixture: a

comparative study." Construction and Building Materials 30: 80-91.

Ismail, I., S. A. Bernal, J. L. Provis, R. San Nicolas, D. G. Brice, A. R. Kilcullen, S.

Hamdan and J. S. van Deventer (2013). "Influence of fly ash on the water and

chloride permeability of alkali-activated slag mortars and concretes." Construction

and Building Materials 48: 1187-1201.

Page 221: Mechanical, Thermal and Structural Investigations of One

References

206

Jalal, M., A. Pouladkhan, O. F. Harandi and D. Jafari (2015). "Comparative study on

effects of Class F fly ash, nano silica and silica fume on properties of high

performance self compacting concrete." Construction and Building Materials 94: 90-

104.

Jankowiak, T. and T. Lodygowski (2005). "Identification of parameters of concrete

damage plasticity constitutive model." Foundations of civil and environmental

engineering 6(1): 53-69.

Jansson, R. (2004). "Measurement of concrete thermal properties at high

temperature." Fib task group 4.

Kaja, A., A. Lazaro and Q. Yu (2018). "Effects of Portland cement on activation

mechanism of class F fly ash geopolymer cured under ambient conditions."

Construction and Building Materials 189: 1113-1123.

Kathirvel, P. and S. R. M. Kaliyaperumal (2016). "Influence of recycled concrete

aggregates on the flexural properties of reinforced alkali activated slag concrete."

Construction and Building Materials 102: 51-58.

Ke, X., S. A. Bernal, N. Ye, J. L. Provis and J. Yang (2015). "One‐Part Geopolymers

Based on Thermally Treated Red Mud/NaOH Blends." Journal of the American

Ceramic Society 98(1): 5-11.

Khale, D. and R. Chaudhary (2007). "Mechanism of geopolymerization and factors

influencing its development: a review." Journal of materials science 42(3): 729-746.

Khalil, M. and E. Merz (1994). "Immobilization of intermediate-level wastes in

geopolymers." Journal of Nuclear Materials 211(2): 141-148.

Khaliq, W. and V. Kodur (2011). "Thermal and mechanical properties of fiber

reinforced high performance self-consolidating concrete at elevated temperatures."

Cement and Concrete Research 41(11): 1112-1122.

Kodur, V. (2000). Spalling in high strength concrete exposed to fire: concerns,

causes, critical parameters and cures. Advanced Technology in Structural

Engineering: 1-9.

Kodur, V. (2014). "Properties of Concrete at Elevated Temperatures." ISRN Civil

Engineering 2014.

Kodur, V., F.-P. Cheng, T.-C. Wang and M. Sultan (2003). "Effect of strength and

fiber reinforcement on fire resistance of high-strength concrete columns." Journal of

Structural Engineering 129(2): 253-259.

Kodur, V. and M. Dwaikat (2008). "A numerical model for predicting the fire

resistance of reinforced concrete beams." Cement and Concrete Composites 30(5):

431-443.

Page 222: Mechanical, Thermal and Structural Investigations of One

References

207

Kodur, V., M. Dwaikat and N. Raut (2009). "Macroscopic FE model for tracing the

fire response of reinforced concrete structures." Engineering Structures 31(10): 2368-

2379.

Kodur, V. and W. Khaliq (2010). "Effect of temperature on thermal properties of

different types of high-strength concrete." Journal of materials in civil engineering

23(6): 793-801.

Kodur, V. and R. Mcgrath (2003). "Fire endurance of high strength concrete

columns." Fire technology 39(1): 73-87.

Kodur, V., T. Wang and F. Cheng (2004). "Predicting the fire resistance behaviour of

high strength concrete columns." Cement and Concrete Composites 26(2): 141-153.

Kodur, V. K. and R. McGrath (2006). "Effect of silica fume and lateral confinement

on fire endurance of high strength concrete columns." Canadian journal of civil

engineering 33(1): 93-102.

Koloušek, D., J. Brus, M. Urbanova, J. Andertova, V. Hulinsky and J. Vorel (2007).

"Preparation, structure and hydrothermal stability of alternative (sodium silicate-free)

geopolymers." Journal of Materials Science 42(22): 9267-9275.

Komnitsas, K. and D. Zaharaki (2007). "Geopolymerisation: A review and prospects

for the minerals industry." Minerals engineering 20(14): 1261-1277.

Kong, D. L. and J. G. Sanjayan (2008). "Damage behaviour of geopolymer

composites exposed to elevated temperatures." Cement and Concrete Composites

30(10): 986-991.

Kong, D. L. and J. G. Sanjayan (2010). "Effect of elevated temperatures on

geopolymer paste, mortar and concrete." Cement and Concrete Research 40(2): 334-

339.

Kong, D. L., J. G. Sanjayan and K. Sagoe-Crentsil (2007). "Comparative

performance of geopolymers made with metakaolin and fly ash after exposure to

elevated temperatures." Cement and Concrete Research 37(12): 1583-1589.

Kroehong, W., T. Sinsiri, C. Jaturapitakkul and P. Chindaprasirt (2011). "Effect of

palm oil fuel ash fineness on the microstructure of blended cement paste."

Construction and Building Materials 25(11): 4095-4104.

Kumar, S., R. Kumar and S. Mehrotra (2010). "Influence of granulated blast furnace

slag on the reaction, structure and properties of fly ash based geopolymer." Journal

of materials science 45(3): 607-615.

Kumaravel, S. and S. Thirugnanasambandam (2013). "Flexural behaviour of

reinforced low calcium fly ash based geopolymer concrete beam." Global Journal of

Research In Engineering.

Page 223: Mechanical, Thermal and Structural Investigations of One

References

208

Kumaravel, S., S. Thirugnanasambandam and C. A. Jeyasehar (2014). "Flexural

Behaviour of Geopolymer Concrete Beams with GGBS." IUP Journal of Structuring

Engineering 7(1): 45-54.

Kusbiantoro, A., M. F. Nuruddin, N. Shafiq and S. A. Qazi (2012). "The effect of

microwave incinerated rice husk ash on the compressive and bond strength of fly ash

based geopolymer concrete." Construction and Building Materials 36: 695-703.

Lee, J. and G. L. Fenves (1998). "A plastic‐damage concrete model for earthquake

analysis of dams." Earthquake engineering & structural dynamics 27(9): 937-956.

Li, C., H. Sun and L. Li (2010). "A review: The comparison between alkali-activated

slag (Si+ Ca) and metakaolin (Si+ Al) cements." Cement and Concrete Research

40(9): 1341-1349.

Li, Z. and S. Liu (2007). "Influence of slag as additive on compressive strength of fly

ash-based geopolymer." Journal of Materials in civil engineering 19(6): 470-474.

Lie, T. and J. Woolerton (1988). "Fire Resistance of Reinforced Concrete Columns:

Test Results, National Research Council of Canada." Institute for Research in

Construction, Internal Report(569).

Liew, Y.-M., C.-Y. Heah, L.-y. Li, N. A. Jaya, M. M. A. B. Abdullah, S. J. Tan and

K. Hussin (2017). "Formation of one-part-mixing geopolymers and geopolymer

ceramics from geopolymer powder." Construction and Building Materials 156: 9-18.

Lloyd, N. and B. V. Rangan (2010). Geopolymer concrete: A review of development

and opportunities. 35th conference on our world in concrete & structures.

Lloyd, N. and V. Rangan (2010). Geopolymer concrete with fly ash. Proceedings of

the Second International Conference on Sustainable Construction Materials and

Technologies, UWM Center for By-Products Utilization.

Log, T. and S. Gustafsson (1995). "Transient plane source (TPS) technique for

measuring thermal transport properties of building materials." Fire and materials

19(1): 43-49.

Lyon, R. E., P. Balaguru, A. Foden, U. Sorathia, J. Davidovits and M. Davidovics

(1997). "Fire‐resistant aluminosilicate composites." Fire and materials 21(2): 67-73.

Majewski, T., J. Bobinski and J. Tejchman (2008). "FE analysis of failure behaviour

of reinforced concrete columns under eccentric compression." Engineering

Structures 30(2): 300-317.

Mansour, M., M. T. Abadlia, K. Bekkour and I. Messaoudene (2010). "Improvement

of rheological behaviour of cement pastes by incorporating metakaolin."

Maranan, G., A. Manalo, B. Benmokrane, W. Karunasena and P. Mendis (2016).

"Behaviour of concentrically loaded geopolymer-concrete circular columns

reinforced longitudinally and transversely with GFRP bars." Engineering Structures

117: 422-436.

Page 224: Mechanical, Thermal and Structural Investigations of One

References

209

McCoy, W. and A. Caldwell (1951). New approach to inhibiting alkali-aggregate

expansion. Journal Proceedings.

McLellan, B. C., R. P. Williams, J. Lay, A. Van Riessen and G. D. Corder (2011).

"Costs and carbon emissions for geopolymer pastes in comparison to ordinary

portland cement." Journal of cleaner production 19(9-10): 1080-1090.

Mendes, A., J. Sanjayan and F. Collins (2008). "Phase transformations and

mechanical strength of OPC/Slag pastes submitted to high temperatures." Materials

and Structures 41(2): 345.

Michikoshi, S. (2008). "A Study on Compressive Strength and Strain Behaviour of

Concrete at High Temperatures-Chapter 1 Introduction." Fire Science and

Technology 27(2): 201-208.

Mijarsh, M., M. M. Johari and Z. Ahmad (2014). "Synthesis of geopolymer from

large amounts of treated palm oil fuel ash: application of the Taguchi method in

investigating the main parameters affecting compressive strength." Construction and

Building Materials 52: 473-481.

Mo, K. H., U. J. Alengaram and M. Z. Jumaat (2016). "Structural performance of

reinforced geopolymer concrete members: A review." Construction and Building

Materials 120: 251-264.

Mo, X., C. Yu and Z. Xu (2003). "Long-term effectiveness and mechanism of LiOH

in inhibiting alkali–silica reaction." Cement and Concrete Research 33(1): 115-119.

Nath, P. and P. Sarker (2011). "Effect of fly ash on the durability properties of high

strength concrete." Procedia Engineering 14(2011): 1149-1149.

Nath, P. and P. Sarker (2012). Geopolymer concrete for ambient curing condition.

Australasian Structural Engineering Conference 2012: The past, present and future of

Structural Engineering, Engineers Australia.

Nath, P. and P. K. Sarker (2014). "Effect of GGBFS on setting, workability and early

strength properties of fly ash geopolymer concrete cured in ambient condition."

Construction and Building Materials 66: 163-171.

Nath, P. and P. K. Sarker (2015). "Use of OPC to improve setting and early strength

properties of low calcium fly ash geopolymer concrete cured at room temperature."

Cement and Concrete Composites 55: 205-214.

Nath, P. and P. K. Sarker (2017). "Flexural strength and elastic modulus of ambient-

cured blended low-calcium fly ash geopolymer concrete." Construction and Building

Materials 130: 22-31.

Nath, S. and S. Kumar (2013). "Influence of iron making slags on strength and

microstructure of fly ash geopolymer." Construction and Building Materials 38: 924-

930.

Page 225: Mechanical, Thermal and Structural Investigations of One

References

210

Nath, S., S. Maitra, S. Mukherjee and S. Kumar (2016). "Microstructural and

morphological evolution of fly ash based geopolymers." Construction and Building

Materials 111: 758-765.

Nayal, R. and H. A. Rasheed (2006). "Tension stiffening model for concrete beams

reinforced with steel and FRP bars." Journal of Materials in Civil Engineering 18(6):

831-841.

Nematollahi, B. and J. Sanjayan (2014). "Effect of different superplasticizers and

activator combinations on workability and strength of fly ash based geopolymer."

Materials & Design 57: 667-672.

Nematollahi, B., J. Sanjayan, J. Qiu and E.-H. Yang (2017). "High ductile behaviour

of a polyethylene fiber-reinforced one-part geopolymer composite: A

micromechanics-based investigation." Archives of Civil and Mechanical Engineering

17(3): 555-563.

Nematollahi, B., J. Sanjayan, J. Qiu and E.-H. Yang (2017). "Micromechanics-based

investigation of a sustainable ambient temperature cured one-part strain hardening

geopolymer composite." Construction and Building Materials 131: 552-563.

Nematollahi, B., J. Sanjayan and F. U. A. Shaikh (2015). "Synthesis of heat and

ambient cured one-part geopolymer mixes with different grades of sodium silicate."

Ceramics International 41(4): 5696-5704.

Němeček, J., P. Padevět, B. Patzak and Z. Bittnar (2005). "Effect of transversal

reinforcement in normal and high strength concrete columns." Materials and

Structures 38(7): 665.

Nguyen, K. T., T. A. Le and K. Lee (2016). "Experimental study on flexural strength

of reinforced geopolymer concrete beams." A-A 8: 150.

Nmecek, J. and Z. Bittnar (2004). "Experimental investigation and numerical

simulation of post-peak behaviour and size effect of reinforced concrete columns."

Materials and Structures 37(3): 161.

Noushini, A., F. Aslani, A. Castel, R. I. Gilbert, B. Uy and S. Foster (2016).

"Compressive stress-strain model for low-calcium fly ash-based geopolymer and

heat-cured Portland cement concrete." Cement and Concrete Composites 73: 136-

146.

O’Connor, S. J. and K. J. MacKenzie (2010). "Synthesis, characterisation and

thermal behaviour of lithium aluminosilicate inorganic polymers." Journal of

materials science 45(14): 3707-3713.

Oh, J. E., P. J. Monteiro, S. S. Jun, S. Choi and S. M. Clark (2010). "The evolution of

strength and crystalline phases for alkali-activated ground blast furnace slag and fly

ash-based geopolymers." Cement and Concrete Research 40(2): 189-196.

Olivia, M. and H. Nikraz (2012). "Properties of fly ash geopolymer concrete

designed by Taguchi method." Materials & Design (1980-2015) 36: 191-198.

Page 226: Mechanical, Thermal and Structural Investigations of One

References

211

Pacheco-Torgal, F., J. Labrincha, C. Leonelli, A. Palomo and P. Chindaprasit (2014).

Handbook of alkali-activated cements, mortars and concretes, Elsevier.

Palomo, A., A. Fernández-Jiménez, G. Kovalchuk, L. Ordoñez and M. Naranjo

(2007). "OPC-fly ash cementitious systems: study of gel binders produced during

alkaline hydration." Journal of Materials Science 42(9): 2958-2966.

Pan, Z. and J. G. Sanjayan (2010). "Stress–strain behaviour and abrupt loss of

stiffness of geopolymer at elevated temperatures." Cement and Concrete Composites

32(9): 657-664.

Pan, Z. and J. G. Sanjayan (2012). "Factors influencing softening temperature and

hot-strength of geopolymers." Cement and Concrete Composites 34(2): 261-264.

Pan, Z., J. G. Sanjayan and F. Collins (2014). "Effect of transient creep on

compressive strength of geopolymer concrete for elevated temperature exposure."

Cement and concrete research 56: 182-189.

Pan, Z., Z. Tao, Y.-F. Cao and R. Wuhrer (2018). "Measurement and prediction of

thermal properties of alkali-activated fly ash/slag binders at elevated temperatures."

Materials and Structures 51(4): 108.

Pan, Z., Z. Tao, Y. Cao, R. Wuhrer and T. Murphy (2017). "Compressive strength

and microstructure of alkali-activated fly ash/slag binders at high temperature."

Cement and Concrete Composites.

Pan, Z., Z. Tao, Y. Cao, R. Wuhrer and T. Murphy (2018). "Compressive strength

and microstructure of alkali-activated fly ash/slag binders at high temperature."

Cement and Concrete Composites 86: 9-18.

Part, W. K., M. Ramli and C. B. Cheah (2015). "An overview on the influence of

various factors on the properties of geopolymer concrete derived from industrial by-

products." Construction and Building Materials 77: 370-395.

Parthiban, K., K. Saravanarajamohan, S. Shobana and A. A. Bhaskar (2013). "Effect

of replacement of slag on the mechanical properties of flyash based geopolymer

concrete." International Journal of Engineering and Technology 5(3): 2555-2559.

Peng, M. X., Z. H. Wang, S. H. Shen and Q. G. Xiao (2015). "Synthesis,

characterization and mechanisms of one-part geopolymeric cement by calcining low-

quality kaolin with alkali." Materials and Structures 48(3): 699-708.

Phoo-ngernkham, T., P. Chindaprasirt, V. Sata, S. Pangdaeng and T. Sinsiri (2013).

"Properties of high calcium fly ash geopolymer pastes with Portland cement as an

additive." International Journal of Minerals, Metallurgy, and Materials 20(2): 214-

220.

Popovics, S. (1973). "A numerical approach to the complete stress-strain curve of

concrete." Cement and concrete research 3(5): 583-599.

Page 227: Mechanical, Thermal and Structural Investigations of One

References

212

Potha Raju, M., K. Srinivasa Rao and P. Raju (2007). "Compressive strength of

heated high-strength concrete." Magazine of Concrete Research 59(2): 79-85.

Provis, J. L., G. C. Lukey and J. S. van Deventer (2005). "Do geopolymers actually

contain nanocrystalline zeolites? A reexamination of existing results." Chemistry of

Materials 17(12): 3075-3085.

Provis, J. L. and J. S. J. Van Deventer (2009). Geopolymers: structures, processing,

properties and industrial applications, Elsevier.

Puertas, F. and A. Fernández-Jiménez (2003). "Mineralogical and microstructural

characterisation of alkali-activated fly ash/slag pastes." Cement and Concrete

composites 25(3): 287-292.

Puertas, F., S. Martınez-Ramırez, S. Alonso and T. Vazquez (2000). "Alkali-

activated fly ash/slag cements: strength behaviour and hydration products." Cement

and Concrete Research 30(10): 1625-1632.

Puligilla, S. and P. Mondal (2013). "Role of slag in microstructural development and

hardening of fly ash-slag geopolymer." Cement and Concrete Research 43: 70-80.

Rahman, M. and P. Sarker (2011). Geopolymer concrete columns under combined

axial load and biaxial bending. Proceedings of the CONCRETE 2011 Conference,

The Concrete Institute of Australia.

Rajendran, M. and N. Soundarapandian (2013). "An Experimental Investigation on

the Flexural Behaviour of Geopolymer Ferrocement Slabs." Journal of Engineering

& Technology 3(2).

Rangan, B. V. (2008). "Fly ash-based geopolymer concrete."

Rangan, B. V., D. Hardjito, S. E. Wallah and D. M. Sumajouw (2005). Studies on fly

ash-based geopolymer concrete. Proceedings of the World Congress Geopolymer,

Saint Quentin, France.

Ranjbar, N., M. Mehrali, U. J. Alengaram, H. S. C. Metselaar and M. Z. Jumaat

(2014). "Compressive strength and microstructural analysis of fly ash/palm oil fuel

ash based geopolymer mortar under elevated temperatures." Construction and

building materials 65: 114-121.

Rattanasak, U. and P. Chindaprasirt (2009). "Influence of NaOH solution on the

synthesis of fly ash geopolymer." Minerals Engineering 22(12): 1073-1078.

Rees, C. A., J. L. Provis, G. C. Lukey and J. S. van Deventer (2007). "Attenuated

total reflectance fourier transform infrared analysis of fly ash geopolymer gel aging."

Langmuir 23(15): 8170-8179.

Rickard, W. D., A. v. Riessen and P. Walls (2010). "Thermal Character of

Geopolymers Synthesized from Class F Fly Ash Containing High Concentrations of

Iron and α‐Quartz." International Journal of Applied Ceramic Technology 7(1): 81-

88.

Page 228: Mechanical, Thermal and Structural Investigations of One

References

213

Rickard, W. D., J. Temuujin and A. van Riessen (2012). "Thermal analysis of

geopolymer pastes synthesised from five fly ashes of variable composition." Journal

of non-crystalline solids 358(15): 1830-1839.

Ryu, G. S., Y. B. Lee, K. T. Koh and Y. S. Chung (2013). "The mechanical

properties of fly ash-based geopolymer concrete with alkaline activators."

Construction and Building Materials 47: 409-418.

Saha, S. and C. Rajasekaran (2017). "Enhancement of the properties of fly ash based

geopolymer paste by incorporating ground granulated blast furnace slag."

Construction and Building Materials 146: 615-620.

Sakulich, A. R., E. Anderson, C. Schauer and M. W. Barsoum (2009). "Mechanical

and microstructural characterization of an alkali-activated slag/limestone fine

aggregate concrete." Construction and Building Materials 23(8): 2951-2957.

Salih, M. A., A. A. A. Ali and N. Farzadnia (2014). "Characterization of mechanical

and microstructural properties of palm oil fuel ash geopolymer cement paste."

Construction and Building Materials 65: 592-603.

Samantasinghar, S. and S. P. Singh (2018). "Effect of synthesis parameters on

compressive strength of fly ash-slag blended geopolymer." Construction and

Building Materials 170: 225-234.

Sarker, P. K. (2009). "Analysis of geopolymer concrete columns." Materials and

structures 42(6): 715-724.

Sarker, P. K., R. Haque and K. V. Ramgolam (2013). "Fracture behaviour of heat

cured fly ash based geopolymer concrete." Materials & Design 44: 580-586.

Sarker, P. K., S. Kelly and Z. Yao (2014). "Effect of fire exposure on cracking,

spalling and residual strength of fly ash geopolymer concrete." Materials & Design

63: 584-592.

Sarker, P. K. and S. Mcbeath (2015). "Fire endurance of steel reinforced fly ash

geopolymer concrete elements." Construction and Building Materials 90: 91-98.

Shaikh, F. and V. Vimonsatit (2015). "Compressive strength of fly‐ash‐based

geopolymer concrete at elevated temperatures." Fire and Materials 39(2): 174-188.

Shehab, H. K., A. S. Eisa and A. M. Wahba (2016). "Mechanical properties of fly

ash based geopolymer concrete with full and partial cement replacement."

Construction and Building Materials 126: 560-565.

Singh, B. and G. Ishwarya (2019). Investigations on Aggregate Reactivity in

Geopolymer Concrete. Recent Advances in Structural Engineering, Volume 1,

Springer: 961-970.

Singh, B., G. Ishwarya, M. Gupta and S. Bhattacharyya (2015). "Geopolymer

concrete: A review of some recent developments." Construction and building

materials 85: 78-90.

Page 229: Mechanical, Thermal and Structural Investigations of One

References

214

Škvára, F., L. Kopecký, J. Nemecek and Z. Bittnar (2006). "Microstructure of

geopolymer materials based on fly ash." Ceramics-Silikaty 50(4): 208-215.

Somna, K., C. Jaturapitakkul, P. Kajitvichyanukul and P. Chindaprasirt (2011).

"NaOH-activated ground fly ash geopolymer cured at ambient temperature." Fuel

90(6): 2118-2124.

Srinivasan, S., A. Karthik and D. Nagan (2014). "An investigation on flexural

behaviour of glass fibre reinforced geopolymer concrete beams." Int. J. Eng. Sci.

Res. Technol 3(4): 1963-1968.

Sturm, P. "Hardening, high-temperature resistance and acid resistance of one-part

geopolymers."

Sturm, P., G. Gluth, H. Brouwers and H.-C. Kühne (2016). "Synthesizing one-part

geopolymers from rice husk ash." Construction and Building Materials 124: 961-966.

Sturm, P., G. Gluth, C. Jäger, H. Brouwers and H.-C. Kühne (2018). "Sulfuric acid

resistance of one-part alkali-activated mortars." Cement and Concrete Research 109:

54-63.

Sturm, P., S. Greiser, G. Gluth, C. Jäger and H. Brouwers (2015). "Degree of

reaction and phase content of silica-based one-part geopolymers investigated using

chemical and NMR spectroscopic methods." Journal of materials science 50(20):

6768-6778.

Sujatha, T., K. Kannapiran and S. Nagan (2012). "Strength assessment of heat cured

geopolymer concrete slender column."

Sukmak, P., S. Horpibulsuk and S.-L. Shen (2013). "Strength development in clay–

fly ash geopolymer." Construction and building Materials 40: 566-574.

Sumajouw, D., D. Hardjito, S. Wallah and B. Rangan (2007). "Fly ash-based

geopolymer concrete: study of slender reinforced columns." Journal of materials

science 42(9): 3124-3130.

Sumajouw, M. and B. V. Rangan (2006). "Low-calcium fly ash-based geopolymer

concrete: reinforced beams and columns."

Suwan, T. and M. Fan (2014). "Influence of OPC replacement and manufacturing

procedures on the properties of self-cured geopolymer." Construction and Building

Materials 73: 551-561.

Swanepoel, J. and C. Strydom (2002). "Utilisation of fly ash in a geopolymeric

material." Applied geochemistry 17(8): 1143-1148.

Taha, B. and G. Nounu (2008). "Using lithium nitrate and pozzolanic glass powder

in concrete as ASR suppressors." Cement and Concrete Composites 30(6): 497-505.

Tao, J., Y. Yuan and L. Taerwe (2010). "Compressive strength of self-compacting

concrete during high-temperature exposure." Journal of Materials in Civil

Engineering 22(10): 1005-1011.

Page 230: Mechanical, Thermal and Structural Investigations of One

References

215

Tao, Z., M. Ghannam, T.-Y. Song and L.-H. Han (2016). "Experimental and

numerical investigation of concrete-filled stainless steel columns exposed to fire."

Journal of Constructional Steel Research 118: 120-134.

Tao, Z. and Q. Yu (2008). "Behaviour of CFRP-strengthened slender square RC

columns." Magazine of Concrete Research 60(7): 523-533.

Temuujin, J., A. van Riessen and K. MacKenzie (2010). "Preparation and

characterisation of fly ash based geopolymer mortars." Construction and Building

Materials 24(10): 1906-1910.

Temuujin, J., R. Williams and A. Van Riessen (2009). "Effect of mechanical

activation of fly ash on the properties of geopolymer cured at ambient temperature."

Journal of materials processing technology 209(12-13): 5276-5280.

Temuujin, J. v., A. Van Riessen and R. Williams (2009). "Influence of calcium

compounds on the mechanical properties of fly ash geopolymer pastes." Journal of

hazardous materials 167(1-3): 82-88.

Thomas, R. J. and S. Peethamparan (2015). "Alkali-activated concrete: Engineering

properties and stress–strain behaviour." Construction and building materials 93: 49-

56.

Thorenfeldt, E. (1987). Mechanical properties of high-strength concrete and

applications in design. Symposium Proceedings, Utilization of High-Strength

Concrete, Norway, 1987.

Ukrainczyk, N. and T. Matusinović (2010). "Thermal properties of hydrating calcium

aluminate cement pastes." Cement and Concrete Research 40(1): 128-136.

van Jaarsveld, J. and J. Van Deventer (1999). "Effect of the alkali metal activator on

the properties of fly ash-based geopolymers." Industrial & Engineering Chemistry

Research 38(10): 3932-3941.

Van Jaarsveld, J., J. Van Deventer and G. Lukey (2002). "The effect of composition

and temperature on the properties of fly ash-and kaolinite-based geopolymers."

Chemical Engineering Journal 89(1-3): 63-73.

Van Jaarsveld, J., J. Van Deventer and A. Schwartzman (1999). "The potential use of

geopolymeric materials to immobilise toxic metals: Part II. Material and leaching

characteristics." Minerals Engineering 12(1): 75-91.

Wallace, K., R. Hill, J. Pembroke, C. Brown and P. Hatton (1999). "Influence of

sodium oxide content on bioactive glass properties." Journal of Materials Science:

Materials in Medicine 10(12): 697-701.

Wallah, S. and B. V. Rangan (2006). "Low-calcium fly ash-based geopolymer

concrete: Long-term properties."

Xu, H., J. L. Provis, J. S. van Deventer and P. V. Krivenko (2008). "Characterization

of aged slag concretes." ACI Materials journal 105(2): 131.

Page 231: Mechanical, Thermal and Structural Investigations of One

References

216

Xu, H. and J. Van Deventer (2000). "The geopolymerisation of alumino-silicate

minerals." International journal of mineral processing 59(3): 247-266.

Xu, H. and J. S. Van Deventer (2002). "Geopolymerisation of multiple minerals."

Minerals engineering 15(12): 1131-1139.

Xu, Y.-y. and B. Wu (2009). "Fire resistance of reinforced concrete columns with L-,

T-, and+-shaped cross-sections." Fire Safety Journal 44(6): 869-880.

Yang, K.-H. and J.-K. Song (2009). "Workability loss and compressive strength

development of cementless mortars activated by combination of sodium silicate and

sodium hydroxide." Journal of materials in Civil Engineering 21(3): 119-127.

Yang, K.-H., J.-K. Song, A. F. Ashour and E.-T. Lee (2008). "Properties of

cementless mortars activated by sodium silicate." Construction and Building

Materials 22(9): 1981-1989.

Yang, T., X. Yao, Z. Zhang and H. Wang (2012). "Mechanical property and structure

of alkali-activated fly ash and slag blends." Journal of Sustainable Cement-Based

Materials 1(4): 167-178.

Yip, C. and J. Van Deventer (2003). "Microanalysis of calcium silicate hydrate gel

formed within a geopolymeric binder." Journal of Materials Science 38(18): 3851-

3860.

Yip, C. K., G. Lukey and J. Van Deventer (2005). "The coexistence of geopolymeric

gel and calcium silicate hydrate at the early stage of alkaline activation." Cement and

concrete research 35(9): 1688-1697.

Yost, J. R., A. Radlińska, S. Ernst, M. Salera and N. J. Martignetti (2013).

"Structural behaviour of alkali activated fly ash concrete. Part 2: structural testing

and experimental findings." Materials and structures 46(3): 449-462.

Yusuf, M. O., M. A. M. Johari, Z. A. Ahmad and M. Maslehuddin (2014). "Strength

and microstructure of alkali-activated binary blended binder containing palm oil fuel

ash and ground blast-furnace slag." Construction and Building Materials 52: 504-

510.

Zhao, R. and J. G. Sanjayan (2011). "Geopolymer and Portland cement concretes in

simulated fire." Magazine of Concrete research 63(3): 163-173.