You are on page 1of 87

Journal Pre-proof

Clean production and properties of geopolymer concrete; A review

Y.H.Mugahed Amran, Rayed Alyousef, Hisham Alabduljabbar, Mohamed El-


Zeadani

PII: S0959-6526(19)34549-4
DOI: https://doi.org/10.1016/j.jclepro.2019.119679
Reference: JCLP 119679

To appear in: Journal of Cleaner Production

Received Date: 22 September 2019


Accepted Date: 11 December 2019

Please cite this article as: Y.H.Mugahed Amran, Rayed Alyousef, Hisham Alabduljabbar, Mohamed
El-Zeadani, Clean production and properties of geopolymer concrete; A review, Journal of Cleaner
Production (2019), https://doi.org/10.1016/j.jclepro.2019.119679

This is a PDF file of an article that has undergone enhancements after acceptance, such as the
addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive
version of record. This version will undergo additional copyediting, typesetting and review before it
is published in its final form, but we are providing this version to give early visibility of the article.
Please note that, during the production process, errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier.


Journal Pre-proof

Clean production and properties of geopolymer concrete; A review

 Author affiliation

 Y. H. Mugahed Amran, PhD


Position: Assistant Professor
Address: Department of Civil Engineering, College of Engineering, Prince Sattam Bin Abdulaziz
University, 11942 Alkharj, Saudi Arabia and in the department of Civil Engineering,
Faculty of Engineering, Amran University, 9677 Quhal, Amran, Yemen.
Email: m.amran@psa.edu.sa and mugahed_amran@hotmail.com
 Rayed Alyousef, PhD
Position: Dean of faculty of engineering
Address: Department of Civil Engineering, College of Engineering, Prince Sattam Bin Abdulaziz
University, 11942 Alkharj, Saudi Arabia.
Email: r.alyousef@psa.edu.sa
 Hisham Alabduljabbar, PhD
Position: Vice dean of technical affairs
Address: Department of Civil Engineering, College of Engineering, Prince Sattam Bin Abdulaziz
University, 11942 Alkharj, Saudi Arabia.
Email: h.alabduljabber@psa.edu.sa
 Mohamed El-Zeadani, PhD Candidate
Position: PhD Student
Address: Department of Civil Engineering, Faculty of Engineering, Universiti Putra Malaysia,
43400 Serdang, Selangor, Malaysia
Email: mohamed.elzeadani9595@gmail.com
Journal Pre-proof

Clean production and properties of geopolymer concrete; A review

Abstract

The incessant production of cement has increased the amount of CO2 being released into the

atmosphere; thus, aggravating the issue of global warming which has an adverse effect on the

environment. Therefore, a more sustainable approach and a careful review of the existing

admixtures used to replace conventional concrete have become highly imperative. To this end,

many investigations on geopolymer concrete (GeoPC), which exhibit similar or better durability

and high strength when compared to conventional concrete, have been carried out by various

researchers. GeoPC concrete has the advantage of cement replacement with supplementary

cementitious materials that are combined with alkali activated solutions. GeoPC is a relatively

new, innovative and sustainable engineering material with many advantages over ordinary

concrete. For example, it exhibits higher early strength, lower natural resource consumption, low

cost and ability to form various structural shapes. GeoPC is an essential material that can be used

for concrete building repairs, maintenance of road transport infrastructure and reducing the

negative environmental effects. Therefore, this paper presents a comprehensive review of GeoPC

material, its constituents, production techniques, curing regimes, properties and its potential

applications in the construction industry.

Keywords: Alkali activation solution, Carbon dioxide emission, Geopolymer Concrete,


Geopolymer, Properties, Supplementary cementitious materials.

1
Journal Pre-proof

Table of Contents

1 Introduction 4
2 Constituent materials of GeoPC 10
2.1 Source of by-product materials 10
2.1.1 Fly ash 12
2.1.2 Silica fume 15
2.1.3 Rice husk ash 18
2.1.4 Red mud 20
2.1.5 Ground granulated blast slag 22
2.1.6 Fine aggregates 24
2.1.7 Coarse aggregates 26
2.2 Alkaline liquids system 29
3 Clean production of GeoPC 32
3.1 Composition of GeoPC 32
3.2 Curing regimes of GeoPC 34
3.2.1 Ambient curing 34
3.2.2 Steam curing 35
3.2.3 Oven curing 36
4 Fresh properties 37
4.1 Stability 37
4.2 Compatibility 38
4.3 Setting time 40
4.4 Workability 41
4.5 Deformability 43
5 Mechanical properties 44
5.1 Compressive strength 45
5.2 Splitting tensile strength 48
5.3 Flexural strength 50
5.4 Modulus of elasticity 52
5.5 Stress–strain behavior 53
5.6 Rate of strength development 55
6 Physical properties 57
6.1 Density 57
6.2 Dry shrinkage 59
6.3 Porosity 61
6.4 Sorptivity 63
7 Conclusion 65
8 Acknowledgment 67
9 References 68

2
Journal Pre-proof

List of Acronyms List of nomenclatures


GeoPC Geopolymer concrete Al Aluminum
POFA Palm oil fuel ash K Potassium
RHA Rice husk ash Na Sodium
RM Red mud Si Silicon
FA Fly ash Po(Ss) Poly(sialates)
SFA Silica fume ash CO2 Carbon dioxide
CDD Concrete digital dilatometer H2O Water
RCA Coarse recycled aggregate Na2SiO3 Sodium silicate
STS Splitting tensile strength NaOH Sodium hydroxide
MoE Modulus of elasticity K2SiO3 Potassium silicate
GGBS Ground-granulated-blast-furnace-slags KOH Potassium hydroxide
C&D Construction and demolition SiO2 Silicon oxide
NCA Natural coarse aggregates Al2O3 Aluminum oxide
ALS Alkaline liquids system CaO Calcium oxide
AAS Alkaline-activator solution MgO MgO oxide
SS/SH Sodium silicate to sodium hydroxide Fe2O3 Iron oxide
SPs Super-plasticizers SO3 Sulfur trioxide
CDD Concrete digital dilatometer S Sulfide Sulfur
DS Dry shrinkage IR Insoluble Residue
AAC Alkali-activated cement Na2O Sodium oxide
LOI Loss in ignition Ca(OH)2 Calcium hydroxid
PFA Pulverized fuel ash Al(OH)3 Aluminium hydroxide
ITZ Interfacial transition zone Na2SiO3 Sodium silicate
OPC Ordinary Portland cement

3
Journal Pre-proof

1 Introduction

Concrete is an important material used in construction industries all over the world (Shaikh,

2016). It is considered to be the most commonly utilized building material due to its considerably

low price, durability, availability of constituent materials and ability to be formed into any shape

or size (Basha S et al., 2016; Shaikh, 2016). The binding techniques and materials used for

producing concrete are also considered essential in construction technology (Lakshmi and

Nagan, 2011). For instance, cement is the most extensively used binding material in plain

concrete and reinforced concrete applications (Basha S et al., 2016). The production of Portland

cement (PC) keeps increasing by 9% annually, worldwide. This rate of increase poses a great

danger to the environment due to the large volume of CO2 being released into the atmosphere

during cement production (Madheswaran et al., 2013). Specifically, the annual greenhouse gas

emissions from PC production are about 1.5 billion tons, or an average of 6% of the total

emissions, from multiple sectors around the world (Fig. 1) (Castel, 2016; Dhakal, 2009;

Madheswaran et al., 2013). The greenhouse effect prevents the reflection of solar radiation back

into space; thereby, keeping regular temperature on the earth surface limited between 15 °C and

18 °C (Shalini et al., 2016). The CO2 concentration in the atmosphere has recently increased by

about 30%, or 467 Mt, of which 8% came from the UK in 2012 (Basha S et al., 2016;

Davidovits, 2002; Joseph and Mathew, 2012; Li and Xu, 2009; Madheswaran et al., 2013;

Shaikh, 2016; Shalini et al., 2016; Soltaninaveh, 2008). These emissions produce a greenhouse

effect, a natural phenomenon that accounts for around 65% of global warming (Bhikshma,

2012). The volume of CO2 emitted during the production of various concrete components is

summarized in Table 1 (A. Castel, 2016) and can be computed using equation (1). Economic

development requires a highly effective application of renewable and non-renewable resources

4
Journal Pre-proof

(Shalini et al., 2016). Sustainable development of a new admixture to replace ordinary concrete

has become increasingly important as the world continues to face serious environmental

degradation (Han et al., 2014; Shalini et al., 2016).

Fig. 1: Production of GeoPC in the Ancient Roman era (Davidovits, 2015)


Table 1: Concrete volumes and target embodied-energy CO2 emissions (A. Castel, 2016)
Quantity Emission factor Emissions
Strength (MPa) Structural element 3 -e 3
(m ) (t CO2 /m ) (t CO2-e)
15 Blinding 589 0.20 119
32 Footings 489 0.24 119
32 Slabs 1948 0.27 533
40 In situ columns and walls 235 0.27 63
40 Precast walls 1067 0.33 351 -1185

A suitable alternative to ordinary concrete is geopolymer. This material has been used in the past

during the ancient Roman Empire (Fig. 2). The Romans were known for their historic

monumental structures, especially their pioneering use of limestone before the advent of cement

(Davidovits, 2015). Geopolymers are mostly produced by burning watershed materials, a

technique developed by David Easton in 2011, to produce sustainable masonry with less cement

5
Journal Pre-proof

using recycled materials (Palm oil fuel ash (POFA), red mud (RM), silica fume (SF), rice husk

ash (RHA), and fly ash (FA)). These materials are known to enhance sustainability of structures

that rely on concrete masonry units. Furthermore, Geopolymerization includes a chemical

reaction of aluminosilicate oxides Al3+ in IV-V fold synchronization with silicates, vassalage

polymeric-Si-O-Al-O-sialate connections similar to the following:

2(Si2O5,Al2O2) + K2(H3SiO4)2 + Ca(H3SiO4)2 → (K2O,CaO)(8SiO2, 2Al2O3, nH2O).

Poly(sialates) (Po(Ss)) are labeled by the subsequent experimentally-based formulation Mn[–

(SiO2)z– AlO2]n. wH2O (Davidovits, 2015, 1999).

Where;

M is a monovalent cation such as potassium (K+) or sodium (Na+),


z is any 1, 2, 3 or ≥ 3.
n is the amount of multiple condensation.

Residential, 6% Road transport, 16% China, 30% EU, 16%


Other HICs, 8% India, 7%
Energy, 41% Industries, 20%
USA, 195 Other MICs, 11
Other transport, 6% Other sectors, 10% Russia, 7% Japan, 4%
20%
6% 19% 11%

7%
10%
7% 4%

6% 6%

41%
16% 16% 30%

A B

Fig. 2: Percentage of CO2 emitted by different sectors (A) and by countries (B) worldwide
(European Commision, 2014)

6
Journal Pre-proof

The PoS are known as ring polymers and chain with Al3+and Si4+ in IV-crinkle harmonization

with O2 and sequences in from air to semi-crystalline. In 1978, Davidovits proposed that the by-

product of the reaction of alkaline liquid with aluminum (Al) and silicon (Si) can be used for

manufacturing geopolymers (Table 2) (Davidovits, 2015, 2002). The patchy in semi-oystalline

3D-Network silico-aluminate systems were known as "geopolymer" by Davidovits because they

resembled plastics which are polymers in the basis of carbon rather than silicon. Davidovits

further categorized geopolymers depending on the Si:AI ratio as: Crystalline PoS with SI:Al=

1:1 ratio, Po(S-disiloxo) Mn-(Si-O-A1-O-Si-O-Si-O)n with SI:Al= 2:1 ratio and Po(S-silcao)

Mn-(Si-O-A1-O-Si-O-)n with SI:Al= 3:1 ratio (Davidovits, 1999).

In addition, Davidovits stated that the atomistic proportion Si:Al in the PoS system controls its

characteristics and utilization. For example, a small ratio of Si:Al (1.23) initiates a 3D-system

that is extremely firm. A great ratio of Si:Al, higher than 1.5, causes greater polymerization. The

Si:Al ≥ 3:1 with Po(S-multisiloxo), the polymeric building obtains from the irritated connecting

of PoS sheets, networks or chains with a PoS link (3D or 2D-Networks)l (Davidovits, 2015,

1999). The subsequent reactions arise during geopolymerisation (Davidovits, 1999).

. .
| |
(Si2O2Al2O2)n + H2O + OH- → Si(OH)4 + Al(OH)4- → (–Si–O–Al–O–) + aH2O
| |
O O 𝑛
Geopolymer precursors refer to a family of inorganic pastes with a typical magnitude of not more

than 20 μm. They have a small loose density between 0.54 and 0.86 g/cm3, a high surface area

from 300 m2/kg to 500 m2/kg, a bright texture and a round shape. These materials also contain

dense spheres, cenospheres, porous unburnt carbon, irregular-shaped debris, and blast furnace

slag (created by reducing iron ore with coke in a blast furnace at temperatures of 1350 °C to

1550 °C). They also have the same chemical compositions as zeolites with amorphous

7
Journal Pre-proof

microstructures (Davidovits, 1999). Geopolymer precursors usually have a glass content in

excess of 95% and are crushed into fine binder (ground-granulated-blast-furnace-slags (GGBS))

manufactured by calcining kaolin at a temperature range between 650 °C and 800 °C) (Yang et

al., 2013). These materials mainly consist of amorphous SiO2 and Al2O3 with great pozzolanic

activity (Khale and Chaudhary, 2007). Apart from its filling effect, calcining kaolin at a high

temperature can also react with calcium hydroxide, the NASH gel, and the coexistance (NASH

and CHS) in the final matrix, a product of PC hydration, to produce calcium silicate hydrate gels.

A combination of these materials, plus other alternative raw materials with silica and alumina

content, can minimize CO2 emissions and reduce the harmful environmental effect of cement

manufacturing (Alam et al., 2014). Thus, the percentage of CO2 emissions can be determined

using Eqn. (1) (Kavitha et al., 2016).


𝐶𝑖 ― 𝐶0
CO2 emission, (100%) = 𝐶0 × 100 (1)

Where, – C0 = CO2 emission of the control mix; and


– Ci = CO2 emission of the blended GeoPC mixes.

In the design of GeoPC, another gel nuclei particle is used which should be stable enough to

resist depolymerization and to start a new gel phase that will be responsible for enhancing the

strength and durability of GeoPCs. Moreover, geopolymeric materials have recently attracted

considerable attention from researchers due to their environmental benefits, such as reduced CO2

emissions and reduced depletion of natural resources (Sashidhar et al., 2015). Also,

geopolymeric materials can be used as an alternative cementitiuos material due to its greater

durability and advanced mechanical properties. Fig. 3 summaries the elementary difference

between GeoPC and ordinary Portland concrete (OPC), and illustrates why GeoPC is more

preferable than OPC, particularly in terms of sustainability and durability. GeoPC is a self-

compacting material; that is, it is capable of consolidating under its own weight and it is

8
Journal Pre-proof

considered as the most revolutionary progress in concrete technology. This material, which is

often used in complex structural formworks, can reduce industrial wastes by at least 12.2 Mt

every year, and can emit 5 to 6 times less CO2 when compared to PC (Anuradha et al., 2014;

Reddy, 2015; Shaikh, 2016). Producing one ton of PC releases approximately one ton of CO2 to

the air as an effect of the decarbonation of limestone in a kiln at the time of the production of this

cement (Chindaprasirt et al., 2014; Dimas et al., 2009; Talakokula et al., 2016). Several

investigators have tried to replace PC with a more environment-friendly concrete that uses

certain byproduct materials which will be comprehensively reviewed in the following

subsections. A paradigm shift to GeoPC has been witnessed in construction industries all over

the world due to its indispensable role in reducing the amount of pollutants and CO2 generated

during PC production.

Type of concrete

OPC GeoPC
Liable to freeze and thaw cycling High resistance to chemical

Restricted techniques to accelerate strength growth Appropriate Resistance to freeze and thaw cycling

Liable to small pH material attack It can resist elevated heat without degradation

Restricted to set times Adjustable remedying techniques for final strength with times

Classic and problematic to change modulus characteristics Quicker or slower set times

Extreme age degradation and carbonation when heated Dynamic and designer nominated modulus characteristics

Certain control on penetrability Provision an insulating R-value equal to cellular material

Restricted characteristics It owned an tremendously stumpy air and water porousness

Fig. 3: Performance and properties comparison between GeoPC and OPC

9
Journal Pre-proof

Therefore, the objective of this review paper is to provide a comprehensive review on the

constituents, clean production techniques, heat curing methods and properties of GeoPC as well

as to comprehensively review the literature to provide insights into the potential application of

GeoPC material in the construction industry today.

2 Constituent materials of GeoPC

Generally, PC is not required in the production process of GeoPC. The two key constituents of

GeoPC are: (1) alkaline solutions such as sodium silicate (Na2SiO3), sodium hydroxide (NaOH),

potassium silicate (K2SiO3), and potassium hydroxide (KOH), and (2) Alumino-silicate sources

of byproduct materials such as RM, SFA, RHA, FA, GGBS and fine and coarse aggregates.

2.1 Source of by-product materials

Geopolymer is a patchy silicate-alumino cementing material that is produced by poly

condensation reaction of alkali and polysilicates geopolymeric precursor known as the

geopolymerization process (Davidovits, 2015; Dimas et al., 2009). Geopolymerization is an

advanced process capable of converting several silicate-alumino materials into valuable products

named inorganic polymers or geopolymers (Davidovits, 1999). Geopolymerization includes an

inhomogeneous chemical reaction between alkali solutions and silicate-alumino oxides at mild

temperatures. It is a highly alkaline conditions squashy patchy to semi-crystalline polymeric

systems that contain Si–O–Si and Si–O–Al bonds (Davidovits, 2015, 1999; Dimas et al., 2009).

In the geopolymerization process, a Si- and Al-rich source material reacts with a highly alkaline

solution to create a binding material like a 3D network of poly-(sialates), amorphous to semi-

crystallinity of geopolymers. The development history of concrete technology reveals that using

various supplementary materials to reduce up to 40% of water–cement ratio can result in subtle

technological changes (Table 2). Fine particles can be introduced into concrete by using

10
Journal Pre-proof

sustainable materials. However, nano-powders have a remarkably large surface area that can

greatly change both surface morphology and surface energy (Fig. 4). By modifying their basic

properties and chemical reactivity (Klabunde and Richards, 2009), these factors can enhance the

catalytic ability of nano-materials (Zhang et al., 1998). The new nano-materials that are based on

metals (lithium-ion, sodium-ion, lithium-ion and sulfur-ion), oxides (SiO2, Al2O3, CaO, MgO,

Fe2O3, SO3, and S) and germanium demonstrate a superplastic behavior and require 100% to

10000% elongation before it encounters failure (Boyd et al., 2002). A GeoPC with ultrahigh

strength and increased durability is expected to be developed in the near future.

Fig. 4: Specific surface area and particle size of different binder materials (Sobolev and
Gutiérrez, 2005)

11
Journal Pre-proof

Table 2: Typical properties of byproduct materials (ACI 233R-95 Committee Report, 1997; B.
N. Sangeetha, 2015; Calderón-Moreno et al., 2002; Kabir et al., 2015; Suresh and Nagaraju,
2015)
Property/ Element FA RHA GGBS SFA POFA
Fineness (m /kg) 4900-5200
450 ~ 450 m2/kg 350 to 550 15,000 to 35,000
cm2/g
Bulk density (kg/m) 1300 96-160 kg/m3 1200 1350-1510 2.40-2.50 g/cm3
Specific gravity 2.2 2.11 2.9 2.2 2.14
Silicon (SiO2) 38 to 55 > 90 30 to 40 > 85 > 80
Aluminum (Al2O3) 20 to 40 >9 5 to 20 <2 16-18
Iron (Fe2O3) 6 to 16 > 2.8 <2 8-10
Calcium (CaO) 1.8 to 10 1-2.2 35 to 40 <1 5-18
Magnesium (MgO) 1 to 5 >1 5 to 18 > 1.2

2.1.1 Fly ash

GeoPC may be formed by using low-calcium FA obtained from coal-combustion binder stations

as a by-product of bituminous or anthracite coal combustion (Adam, 2009; American Concrete

Institute. and Malhotra, 2000; Talakokula et al., 2016). FA is identified as “pulverised fuel ash,”

and it is a byproduct of coal combustion which comprises fine particles that have been blown out

of the boiler along with flue gases (flue gas refers to the burning exhaust gas formed at power

plants) (Davidovits, 1999; Liu et al., 2016). FA is often used as a substitute to ordinary PC

(OPC) in concrete production (Anuradha et al., 2014; Castel and Foster, 2015; He et al., 2012;

Kovacik et al., 2011; Liu et al., 2016; Talakokula et al., 2016; Ukwattage et al., 2013; Yildiz,

2004; Zhuang et al., 2016a). Compared with traditional concrete, FA concrete has a higher

strength and durability (Anuradha et al., 2014), can be poured readily, has lesser permeability,

and resists the alkali–silica reaction more efficiently; thereby, extending its service life and

lowering its cost (Zhuang et al., 2016b). For example, the low-calcium FA has been positively

used to produce GeoPC when the aluminum and silicon oxides constitute about 80% by weight,

with the Si:Al ratio of 2 (Adam, 2009; American Concrete Institute. and Malhotra, 2000) . The

iron oxide content is typically varied between 10 and 20% by weight, while the calcium oxide

content is lower than 5% by weight (Adam, 2009). The FA carbon content, as designated by the

12
Journal Pre-proof

loss on ignition by weight, is below 2%. Reportedly, the distribution of FA particle size is about

80% of the FA particles which are less than 50 μm (Adam, 2009; Castel and Foster, 2015; Law

et al., 2015; Talakokula et al., 2016). FA can be 20% to 60% cheaper than OPC in some

countries, but in some cases, OPC can be more than twice as expensive as FA (Chindaprasirt et

al., 2014; Kovacik et al., 2011; Kumar and Kumar, 2013; Liu et al., 2016; Shaikh, 2016; Zhuang

et al., 2016a). However, FA is rarely shipped at long distances and is more expensive than local

OPC because some concrete durability requirements can only be fulfilled by using FA. This

material can also positively affect the environment due to the conservation of the landfill spaces,

reduction of water and energy consumption, and minimization of greenhouse gas emissions

(Anuradha et al., 2014; Chang and Shih, 2000; Chindaprasirt et al., 2014; Talakokula et al.,

2016; Ukwattage et al., 2013; Yildiz, 2004; Zhuang et al., 2016b). In this case, FA can reduce the

production of OPC that emits about 1 ton of carbon dioxide for every ton of cement

manufactured; in other words, for each ton of FA used, the CO2 emissions are cut by one ton

(Castel and Foster, 2015; Talakokula et al., 2016; Zhuang et al., 2016b). In fact, using an entire

year’s supply of FA for concrete production is equivalent to eliminating 25% of the CO2 emitted

by vehicles worldwide (Poudenx, 2008). The fineness of FA is calculated based on ASTM C115

standard (ASTMC115-96, 1996) as illustrated in Fig. 5 (Sanjayan et al., 2015). FA particles’

average size and blain surface area are usually 9 μm and 0.37 m2/g, respectively. Given the very

small size of its particles, FA can enhance the density and frost resistance of GeoPC and reduce

its permeability (Table 3) (ASTMC666, 1997).

13
Journal Pre-proof

Fig. 5: Approximate particle size distribution of FA (Li et al., 2014)

Table 3: Frost resistance of FA (ASTMC666, 1997)


Results at 300 cycles
Fly ash mixtures Frost resistance in water, ASTM C 666 Method A [AASHTO T 161]
Expansion, % Mass loss, % Durability factor
Average of: Class C, 0.006 1.6 101
Class F 0.004 1.8 102
Control mixture 0.002 2.5 101

FA production has become significant because of its key role in the economic and green

utilization of technologies (Fig. 6). This material can also be used in soil amendment (Ukwattage

et al., 2013), nutrient retrieval (Kovacik et al., 2011), waste removal (as a low-cost absorbent)

(RUBEL et al., 2005; Yildiz, 2004), and zeolite production (as a source of Si and Al) (Chang and

Shih, 2000; Lloyd et al., 2012). FA has been adopted recently as an alternate material for

producing geopolymers, a new cement or binder that is similar to hydrated cement in terms of

reactivity, appearance, and other characteristics (Zhuang et al., 2016b). The chemical

compositions of OPC and FA are given in Table 4.

14
Journal Pre-proof

Table 4: Chemical compositions of OPC and FA (Malathy, 2009)


Comp. LOI IR SiO2 Al2O3 Fe2O3 CaO Mgo So3 Na2O Specific
surface
Cement 1.65 0.76 19.33 5.66 2.66 63.07 0.36 3.38 0.14 4.51 m²/kg
Fly Ash 2.88 0.30 54.92 23.04 6.62 3.84 2.82 0.76 0.73 3.60 m²/kg

Fig. 6: Application fields of FA in the U.S. (Adams, 2017)


2.1.2 Silica fume

The Silica Fume Association was established in 1998 to help silica fume manufacturers to

promote the application of silica fume in concrete (Anuradha et al., 2014; Saraya, 2014). As a

byproduct of Si metal or ferrosilicon alloy production (highly-reactive pozzolan and a

fundamental ingredient in high-performance concrete), silica fume ash (SFA) was used to

replace 10 – 40 % of OPC content in concrete (Tolêdo Filho et al., 2003). Based on its physical,

chemical and mineralogical properties, SFA is a highly reactive pozzolan that may come from

either natural or artificial sources (Fig. 7) (Sreenivasulu, C. and Jawahar, 2015; Triantafillou,

2016). The silica in pozzolana reacts with the portlandite formed during the hydration of OPC

and assists in its strength development (Singh et al., 2015; Sreenivasulu, C. and Jawahar, 2015).

15
Journal Pre-proof

Furthermore, this material progressively creates calcium silicate hydrate, a binder that takes up

the space in concrete materials and enhances their impermeability, durability, and strength

(Sakulich, 2011; Sashidhar et al., 2016; Singh et al., 2015). The hydration of OPC can be written

as

C3S + H2O ➟ C-S-H + Ca(OH)2,

where

‒ C-S-H = calcium–silicate–hydrate; and


‒ Ca(OH)2 = Portlandite.
Meanwhile, the combination of limestone with the silica of pozzolana can be expressed as:

Ca(OH)2+SiO2 ➟ C-S-H

where

‒ SiO2 = Silica

Fig. 7: Artificial production of SFA (Naik, 2008)

Amorphous silica has a faster reaction to silica compared with crystalline silica. Such difference

accounts for the variation among active pozzolanas and materials of comparable chemical

16
Journal Pre-proof

composition that display minimal pozzolanic activity (Kovler and Roussel, 2011; Liu et al.,

2016; Nematollahi et al., 2017). However, concrete-containing SFA can demonstrate very high

strength and durability (Geopolymer and 1988, n.d.; Kovler and Roussel, 2011). SFA can be

obtained from concrete admixture suppliers and, when stated, is easily added during the

production of concrete (with a limit between 10 and 20 %) (American Concrete Institute. and

Malhotra, 2000; Karbhari, 2013; Subang Jaya et al., 2013). The placement, finishing and curing

of SFA-based GeoPC require special attention from concrete contractors (Geopolymer and 1988,

n.d.; Poon et al., 2006). The smoke that is emitted during furnace operations is collected and sold

as SFA instead of being landfilled (Al-Qadri et al., 2009). SFA can also be included in concrete

as a mineral admixture (Al-Qadri et al., 2009; Bhavsar et al., 2014; MATSAGAR, 2015; Mo et

al., 2016). Silica fume principally consists of amorphous (non-crystalline) silicon dioxide (SiO2)

(Siddique and Iqbal Khan, 2011) and extremely small particles that are approximately 1/100 the

size of an average cement particle (M. Greim and W. Kusterle, 2004). Given its high SiO2

content, large surface area and fine particles, SFA is a highly reactive pozzolan applied in

concrete (Al-Qadri et al., 2009; Detwiler et al., 1996; Siddique and Iqbal Khan, 2011;

Triantafillou, 2016). The quality of SFA is specified by ASTM C 1240 (ASTM, 2012) and

AASHTO M 307. Concrete structures in the US often deteriorate by the corrosion induced by

deicing or marine salts (Wolsiefer, 1991). SFA concrete with low water content is greatly

defiant to infiltration by chloride ions (Neville, 1987). Given its desirable properties,

transportation agencies have begun to use SFA instead of concrete for constructing new bridges

or rehabilitating existing structures (Sakulich, 2011). The specifications of silica fume concrete

with high durability or strength can be acquired from the suppliers of SFA or other key

admixtures (Anuradha et al., 2014; ASTM, 2012; Bhavsar et al., 2014; Poon et al., 2006; Saraya,

17
Journal Pre-proof

2014). Silica fume can be added in its wet or dry form during the production of concrete at a

manufacturing plant (Atiş et al., 2005). SF has been effectively manufactured in both dry batch

plants and central mix as an alternative to concrete (Kamath and Khan, 2016). Several guidelines

have also been issued concerning the handling and use of silica fume for producing high-quality

concrete (Mazloom et al., 2004).

2.1.3 Rice husk ash

Rice husk ash (RHA) is a carbon-neutral green product (Table 5) (Junaid et al., 2014; Mehta,

1977) that is mostly used as ash for generating power (Ajay et al., 2012), or as boiler fuel for

processing paddy with volume between 20% and 25% of the rice paddy is an indigestible outer

husk that is removed and burnt either in household stoves or in local power plants to produce

steam for boiling rice (Sanjayan et al., 2015). In contrast, in the RHA-based GeoPC structures,

the presence of sodium silicates with alkali-activated at elevating temperature can lead to

producing new crystalline phases, for example Na-feldspars, albite (NaAlSi3O8), and nepheline

(NaAlSiO4), contributing to a higher thermal stability at elevated temperatures (Saravanan and

Sivaraja, 2016). The crystalline silica content of RHA has received wide concern because of the

potential hazards of inhaling this mineral (Sanjayan et al., 2015). RHA is approximately 25% by

weight of rice husk when burnt in boilers (Fig. 8) (Raheem et al., 2013). RHA is an excellent

super pozzolan that can be utilized to produce mixes of special concrete (Saravanan and Sivaraja,

2016). This material may be applied as an alternative for cement in concrete production (Habeeb

and Mahmud, 2010; He et al., 2013; Junaid et al., 2014; Mehta, 1977; Raheem et al., 2013;

Saravanan and Sivaraja, 2016; Shalini et al., 2016). Fine amorphous silica is increasingly being

utilized in the manufacturing of special cement and low-permeability, high-strength and high-

performance concrete mixes as well as in the construction of nuclear power plants, marine

18
Journal Pre-proof

environments and bridges (Board, 2012; Sanjayan et al., 2015). Most of the SFA or micro silica

being sold in the market has been imported from Burma, China and Norway (Naji et al., 2010).

In addition, due to supply shortage, the price of silica fumes in India increased to almost US

$500 per ton, which is far greater than the selling prices in China, Canada, and the US (Habeeb

and Mahmud, 2010; Shalini et al., 2016). The cement demand in India is also expected to reach

approximately 550 Mt by 2020 with a shortfall of approximately 230 Mt (–58%), and such

increasing demand can be ascribed to the increasing number of infrastructural activities being

conducted within the country (Singh et al., 2015). Given that RHA comprises about 18% rice

husks, producing one ton of rice will generate approximately 45 kg of RHA, which has

substantial pozzolanic properties, rich silica content (~95%), and high surface area (Gonçalves

and Bergmann, 2007). The amount, crystalline content and chemical composition of the

produced silica strongly depend on furnace design and the burning temperatures (Zain et al.,

2011). In GeoPC, it is reported that RHA blended concrete could lower the temperature influence

that arises during the hydration of cement (B. N. Sangeetha, 2015; Shalini et al., 2016). RHA

blended concrete can improve the workability of concrete in comparison with OPC and could

also raise the setting time of cement pastes (Detphan and Chindaprasirt, 2009; Mehta, 1977;

Nazari et al., 2011). Moreover, RHA-based GeoPC can reduce concrete’s total porosity, adjust

its pore configuration, and considerably lower the permeability which permits the effect of

dangerous ions contributing to the weakening of the concrete matrix (Habeeb and Mahmud,

2010; He et al., 2013; Naji et al., 2010). RHA cement increases the compressive strength and

assists in improving the initial age mechanical and long-term strength characteristics of GeoPC

pastes. Specifically, partial substitution of cement with RHA decreases the water diffusion into

19
Journal Pre-proof

concrete by capillary action and effectively enhances the resistance of GeoPC to sulphate attack

(Ajay et al., 2012; Saravanan and Sivaraja, 2016).

Fig. 8: Synthesis of RHA from rice husks (Gursel et al., 2016)


Table 5: Technical specifications of RHA (Junaid et al., 2014; Shalini et al., 2016; Zain et al.,
2011)
Technical specifications Percentage/ Size/ Value
P 75 % minimum
Humidity 2 % maximum
Particle size 25 microns (d50)
Colour Grey
Loss on ignition at 800°C 4 % maximum
pH value 8

2.1.4 Red mud

Red mud (RM) is an offshoot of the Bayer process (Fig. 8) for refining bauxite to alumina with

volume between 55-65% of the processed bauxite (He et al., 2013), and the resulting alumina

serves as raw material for creating Al via the Hall–Héroult process (Ye et al., 2016, 2014). An

20
Journal Pre-proof

ordinary alumina plant generates one to two times as much RM as alumina. RM consists of 30-

60% iron oxides which account for its red color (He et al., 2013; Si et al., 2013). More

specifically, Dharmendra et al. (Dharmendra S. Ravat and Dave, 2017) reported a value of

42.25% for iron oxide content in RM. Furthermore, this material is extremely basic and has

a pH value of 10 to 13 (Ye et al., 2016). About 2 to 3 Mt of RM are being used annually in

cement production [100], red sand used in the sub-grade and sub-base in road construction

(Biswas and Cooling, 2013), and iron production (Liu et al., 2009; PARAMGURU et al.,

2004). This material can also be used in producing low-cost concrete, studying sandy soils,

improving phosphorus cycles, reducing soil acidity, capping landfills, and sequestering carbon

(Si et al., 2013). The solid constituents of RM mainly consist of alumina, iron oxides (mostly

hematite), and some toxic heavy metals (PARAMGURU et al., 2004). This material also

becomes somewhat radioactive if the original bauxite contains radioactive minerals (Kumar and

Kumar, 2013; PARAMGURU et al., 2004; Si et al., 2013; Ye et al., 2016, 2014). Given the high

alkalinity and water content of this material, the safe and economic disposal of RM poses a

major environmental concern for alumina refineries (Dharmendra S. Ravat and Dave, 2017; He

et al., 2013, 2012; Kumar and Kumar, 2013; PARAMGURU et al., 2004). Table 6 displays the

chemical composition of RM. Reportedly, the addition of RM improved the intensity of the

reaction and structural restructuring; however, enhancement in both compressive strength and

setting time are found when the specimens only possess 5-20% RM, and greater than that could

cause adverse effects on the relative properties of GeoPC (Dharmendra S. Ravat and Dave,

2017; He et al., 2013, 2012; Kumar and Kumar, 2013). Henceforth, it may be deduced that the

low addition of RM in GeoPC as a partial cement substitute can lead to higher solution and

utilization of sodium trisilicate which can have a positive effect on the Young’s modulus, failure

21
Journal Pre-proof

strain and compressive strength of RM-based geopolymers, leading to more reactive Si to the

GeoPC synthesis, resulting in an important contribution to facilitating the geopolymerization

reaction.

Table 6: Chemical composition of RM (Dharmendra S. Ravat and Dave, 2017; He et al., 2013,
2012; Kumar and Kumar, 2013; Ye et al., 2014)
Oxides Approximated value, %
Silicon Dioxide, SiO2 ~ 13.14
Aluminium Oxide, AL2O3 ~ 20.26
Iron Oxide, Fe2O3 ~ 42.25
Calcium Oxide, Cao ~ 1.25
Sodium Oxide, Na2O ~ 4.36
Titanium Oxide, TiO2 ~ 1.9
Finesse ~ 2200 cm2/gm
Specific Gravity ~ 2.09

2.1.5 Ground granulated blast slag

Being a by-product of blast furnaces, Ground granulated blast slag (GGBS) is often utilized in

iron production (Li and Zhao, 2003; Suresh and Nagaraju, 2015). This material can be obtained

at temperatures of almost 1500 °C and is fed with a cautiously controlled mixture of iron ore,

coke, and limestone (Fig. 9) (Suresh and Nagaraju, 2015). The melted slag has a content of about

40% calcium oxide (CaO) and 30-40% silicon dioxide (SiO2), which is near to the chemical

formation of OPC. When iron ore is reduced to iron, the residual materials create a slag that

floats on top of the iron (Ananthayya and WP., 2014; Suresh and Nagaraju, 2015). This slag is

regularly tapped off as a molten liquid and must be quenched quickly in large volumes of water

to manufacture GGBS (Sahithi and Priyanka., 2015). This quenching process utilizes the

cementitious properties of the slag and engenders granules comparable to coarse sand

(Sumajouw et al., 2004). The granulated slag is subsequently desiccated and ground to fine

powder. GGBS can replace the OPC content of concrete by 35% to 70%(Kelham, 1996) and

exhibits excellent properties when finely ground and combined with other materials to form

22
Journal Pre-proof

GeoPC (ACI 233R-95 Committee Report, 1997; Jain and Pal, 1998; Shukla et al., 2009). The

glass particles of GGBS comprise Q0-type mono-silicates, which are comparable to those being

applied in ordinary Portland cement clinker and dissolve upon activation by any medium

(Ganesh Babu and Sree Rama Kumar, 2000; “Prediction of Long-Term Corrosion Resistance of

Plain and Blended Cement concretes,” 1993). GGBS’s glass content typically exceeds 85% of its

total volume, while its specific gravity ranges from 2.7 to 2.90 (lower than that of ordinary

Portland cement), and its bulk density varies between 1200 kg/m3 and 1300 kg/m3 (Saraya,

2014). The standard chemical composition of this material is presented in Table 7. The chemical

composition of ordinary cement shows more similarities to that of GGBS than of other mineral

admixtures, such as POFA concrete (Saraya, 2014). GGBS can be used for refining the pores and

increasing the long-term strength, sulfate, and alkali silica reaction resistance of concrete as well

as for reducing the water demand, permeability, and heat generation during the hydration process

(Castel and Foster, 2015; Dimas et al., 2009; Hadjsadok et al., 2012; Saraya, 2014; Suresh and

Nagaraju, 2015). However, the addition of GGBS can influence the reaction, characteristics and

GeoPC matrix. The influence is varied in the basis of the volume of GGBS added (5–50%). It is

found that the reaction at 27°C is governed by the GGBS activation. The reaction at 27ºC is

contributed by precipitation and dissolution of C–S–H gel due to the alkali activation of GGBS

(Ananthayya and WP., 2014; Li and Zhao, 2003; Sahithi and Priyanka., 2015). Moreover, in the

production of GeoPC, the aluminum and silicon present in the GGBS are activated by a mixture

of sodium silicate and sodium hydroxide solutions to produce the geopolymer paste that binds

the aggregates (Islam et al., 2015, 2014). It may be deduced that the increased addition of GGBS

can increase the ultrasonic pulse velocity, resistance to acid and compressive strength of GeoPC

at all curing regimes(B. N. Sangeetha, 2015; Ganesh Babu and Sree Rama Kumar, 2000).

23
Journal Pre-proof

Fig. 9: Sources of GGBS (Spanlang et al., 2016)

Table 7: Chemical composition of GGBS


Quantities specified Quantities specified by ASTM Annotation
Oxides Practically, % (Saraya, 2014; C989, % (Deb et al., 2015)
Shukla et al., 2009)
Silicon Dioxide, SiO2 30 - 35 40.0 -
Aluminium Oxide, 8 - 22 13.5 Higher Slag
Al2O3
Calcium Oxide, CaO 27 - 32 39.2 Lower Slag
Manganese Oxide MgO 7-9 3.6 -
Iron Oxide, Fe2O3 8 - 10 1.8 Higher Slag
Sulfur ion, SO3 3-7 4 -
Sulfide sulfur, S - 1.0-1.9 -
LOI - 0.0 -
2.1.6 Fine aggregates

Fine aggregates contribute to the production of GeoPC by 35% to 45% (Shaikh, 2016; Shalini et

al., 2016). Foundries have successfully recycled and reused sand many times in the past. This

study examines the efficacy of foundry sand as a partial substitute (up to 25%) for fine

aggregates in GeoPC (Joseph and Mathew, 2012). Self-compacted GeoPC is a composite

material made from a geopolymeric paste, fine and coarse aggregates. Fine aggregates are

24
Journal Pre-proof

usually mined from sand quarries, and red mud is a result of the Bayer process for manufacturing

alumina from bauxite (Fig. 10) (Soltaninaveh, 2008). Electronic waste sand can also be applied

as a partial substitute for fine aggregates in concrete. A concrete mix that contains sand shows

much better properties than conventional concrete because of the even particle size, high reactive

silica content, and high density of sand. Granite powder is a fine aggregate in GeoPC that is

generated from the waste materials of the granite industry (e.g., solid waste and stone slurry),

which annual volumes can reach as high as 12.2 million tons (Reddy, 2015). Table 8 gives the

typical properties of fine aggregates. Meanwhile, it is reported that the appropriate choice for the

ratio of fine aggregate to total aggregate content (60-75%) in GeoPC is 0.35, and this leads to

improvements in the Poisson’s ratio, modulus of elasticity, split and flexural tensile strength of

GeoPC by 19.2%, 14.4%, 45.5% and 30.6%, respectively, compared to normal concrete (Joseph

and Mathew, 2012). Also, researchers reported that the pozzolanic material for higher concrete

strength indicated that optimum performance is attained by replacing 7% to 15% of Metakolin;

however, the increase in FA content led to decrease in strength (Basha S et al., 2016;

Soltaninaveh, 2008). Specifically, it can be inferred that the mechanical properties of GeoPC can

improve when the partial replacement of cement does not exceed 20%, any more than that will

lead to a slight reduction in the concrete structure compared to nominal concrete design mix.

Table 8: Properties of fine aggregates (different types of sands) (Soltaninaveh, 2008)


Characteristic Value
Type Uncrushed (shape)
Specific gravity 2.68
Fineness modulus 2.50
Water adsorption 1.02
Grading zone II

25
Journal Pre-proof

Fig. 10: Bayer process (Dodoo-Arhin et al., 2017)

2.1.7 Coarse aggregates

About 75% of coarse recycled aggregate (RCA) is made from concrete while the rest is made

from masonry, asphalt, tile and others (Shaikh, 2016). RCAs can partially replace natural coarse

aggregates (NCA) in GeoPC by 15%, 30%, and 50% (KThu and Murthy, 2015; Shaikh et al.,

2015). Given the increasing use of concrete every year, the extraction of NCA for producing

concrete can adversely affect the natural ecosystem (KThu and Murthy, 2015; Shaikh, 2016).

The dumping of construction and demolition (C&D) wastes poses another environmental

concern that has inspired several researchers to explore new methods of recycling such wastes

with an aim of increasing the available landfill space and reducing the present dependence on

minerals and natural aggregates (Kou et al., 2012; Zaharieva et al., 2003). Knowing that fine and

coarse aggregates comprise roughly 75% to 80% concrete [9, 121], C&D wastes may be used in

the form of RCA to produce concrete (Corinaldesi and Moriconi, 2009). Although this idea is not

exactly new, many researchers have investigated the properties of RCA-containing concrete and

agreed that such properties are inferior to those of NCA-containing concrete (Corinaldesi and

26
Journal Pre-proof

Moriconi, 2009; Etxeberria et al., 2007; Shaikh, 2014; Shaikh et al., 2015). However, only few

researchers have tested the durability and mechanical properties of RCA-containing GeoPC

(Anuar et al., 2011; Nuaklong et al., 2016; Posi et al., 2013; Sata et al., 2013; Shi et al., 2012).

These studies have also tested the compressive strength of RCA-containing GeoPC that uses

geopolymer from wastepaper sludge ash instead of that from FA and slag. They found that the

compressive strength of this concrete rises by almost 10% from 7 to 28 days and that the high

molarity of sodium hydroxide displays a higher compressive strength in GeoPC than in

conventional concrete. Shuang et al. (Shi et al., 2012) examined the mechanical properties of

GeoPC that contains 50% and 100% RCA as a substitute for NCA, and compared such properties

with those of ordinary concrete. The coarse aggregate used in GeoPC comprises pea gravels with

a maximum size of 10 mm (dmax) (Fig. 11), a bulk dry specific gravity and absorption (ASTM

C127) (ASTMC, 2004) (Table 9). The particle size distribution of the alumino-silicate source by-

product can influence the compressive strength of GeoPC as given in Table 10.

Table 9: Physical properties of coarse aggregates (A. Castel, 2016; Castel and Foster, 2015;
Ganesh et al., 2016; Shrivastava and Shrivastava., 2015)
Property Approximated value
Specific Gravity 2.5 - 2.80
Bulk density 1550 -1570 kg/m3
Compaction Factor 0.85-0.95
Flakiness Index 10.23 - 12.2%
Elongation Index ~ 28.72%
absorption of the aggregates ~ 1%

27
Journal Pre-proof

Fig. 11: Coarse aggregate particle distribution (Adewuyi et al., 2017)


Table 10: Influence of the particle size distribution of the binder phase on the compressive
strength of GeoPC based on the type of poly(sialate) structure
Types
Compressive
of Fineness Major findings Ref.
strength, MPa
GeoPC
RHA- BA: 15.7, 24.5, and 32.2 Finer BA gives rise to higher (Nuaklong et al.,
35–61.5
based µm strength 2016)
FA with Blaine fineness FA with highest Blaine
(Chindaprasirt et al.,
39–75 of 2700, 3900, and 4500 fineness give rise to
2014)
FA- cm2/g optimum strength
based Combination of FFA and
FA: 75 and 3 µm RHA:
15–45 RHBA give rise to (Ye et al., 2016)
90 and 7 µm
highest strength
RHA: 5%, 3%, and 1% Finer RHA gives rise to
(Dharmendra S. Ravat
34.5–43.0 retained on No. 325 highest compressive
and Dave, 2017)
RHA- sieve strength
based 37.43% strength increment
RHA: 100% passes 150
15-16 as compared (Board, 2012)
µm sieve
to unground samples

28
Journal Pre-proof

2.2 Alkaline liquids system

The alkaline liquids system (ALS) is an alkaline-activator solution (AAS) for GeoPC such as

Na2CO3, NaOH, Na2SO4 and Na2O·nSi2 (Table 11). These alkaline activators can be applied in

solid or liquid state. Regularly, cements integrated with activator and precursors are favored (in

solid state) and water is used as a mixing liquid. The gels matrix shaped by slag activation

powerfully relies on numerous chemical factors governing the reaction mechanism and,

consequently, the improvement of durability properties and resistance to external attacks. Also,

ALS is obtained by combining the solutions of alkali silicates and hydroxides, except for

distilled water (Shalini et al., 2016). The alkaline liquids include concentrated aqueous alkali

hydroxide or silicate solution with soluble alkali metals that are commonly potassium (K) or

sodium (Na) based used in the preparation of alkaline activators for balancing the negative

charge of the alumina in four fold coordination with the silica (Madheswaran et al., 2013). These

liquids produce a geopolymeric binder by extracting and activating Si and Al atoms from Si- and

Al-rich by-product materials (Madheswaran et al., 2013). In GeoPC, a large number of research

focused on FA and GGBS as precursor of their blended cements, because they are replete with Si

and Al, and they are activated by alkaline liquids to make the geopolymeric binder (Liang et al.,

2019; Puertas et al., 2003; Ryu et al., 2013). Though, the nature of FA and GGBS (their chemical

and mineral compositions) disturbs the final performance of the GeoPC gained. Also, the key of

AAS is to melt the reactive portion of source materials Si and Al found in FA and provide a high

alkaline liquid medium for concentrated polymerization reaction (Temuujin et al., 2011; Yusuf et

al., 2015). Furthermore, geopolymers are amorphous to semi crystalline polymeric items made

by the alkali activation of alumino-silicate constituents with alkaline silicate solution at ambient

or higher temperatures, leading to possible use as a substitute to ordinary cement (Usha et al.,

29
Journal Pre-proof

2016). However, the effect of the chemical activation process on the compressive strength of

GeoPC is shown in Table 12.

Table 11: Typical alkaline liquids, acids, bases, and their pH scale (Shalini et al., 2016)
pH value H+ concentration relative to pure water Typical alkaline liquids
11 0.000 1 Ammonia solution
12 0.000 01 Soapy water
13 0.000 001 Bleach, oven cleaner
14 0.000 000 1 Liquid drain cleaner

Table 12: Influence of the chemical activation process on the compressive strength of GeoPC
Chemical activator
Types of Compressive
NaOH SS/SH Major findings Ref.
GeoPC Ms Strength, MPa
molarity ratio
7–32 SS/SH ratio of 2.5 optimal (Yan et al., 2016)
POFA-based 0.92–1.63 10 0.5–3.0 Ms of 0.915 optimal but
65–69 (Nazari et al., 2011)
insignificant
(Dharmendra S.
14 and 1.9–5.5 15–40 SS/SH ratio of 4.0 optimal Ravat and Dave,
18 2017)
RHA-based
- 0.5–2.5 34–56 18 M NaOH optimal (Yang et al., 2013)
2–6 8–15 2 M NaOH optimal (Board, 2012)
2.5
4–12 20–30 12 M NaOH optimal (Ye et al., 2016)
HMT-based 10–15 4–34 15 M NaOH optimal (Sivaraja et al., 2010)
0.75–1.25 10 39–57.3 Ms of 1.0 optimal (Kelham, 1996)
-
(Sashidhar et al.,
1.0–2.0 5.0–63.4 Ms of 1.5 optimal
2016)
-
7.5– SS/SH ratio of 1.5 and 7.5 (Chindaprasirt et al.,
25–45
12.5 M NaOH optimal 2013)
-
4.5–
- 4–14 SS/SH ratio of 0.7 optimal (Zhang et al., 2016)
FA-based 16.5
Strength increased from
0.33– 4.5 to 14 M NaOH, but (San Nicolas et al.,
10 7–25
3.0 decreased at 16.5 M 2013)
NaOH
(Nematollahi and
- 3–9 0.4–2.3 12–23 6 M NaOH optimal
Sanjayan, 2014)

Fig. 12 illustrates how the viscosities of alkali hydroxide solutions change along with their

concentration at 25 °C. The datasets lack information for some cases because of the

unavailability of data in the literature (Provis and Deventer, 2009). However, all solutions show

a consistent trend. Specifically, their viscosity gradually increases up to 1.0 M, after which the

viscosity no longer shows any significant change from that of water, meaning that the viscosity

30
Journal Pre-proof

of a solution varies with a rise in the molarity of the solution (Usha et al., 2016). Further, the

viscosity significantly increases after this point, and the steepness of such increment depends on

the identity of the alkali cation. Fig. 12 is plotted on logarithmic axes. When treated as a function

of cation size, viscosity does not show any systematic trend but demonstrates a significant degree

of uncertainty given the lack of data for some systems (Provis and Deventer, 2009).

Fig. 12: Viscosities of alkali hydroxide solutions as a function of molarity (Olsson et al., 1997;
Provis and Deventer, 2009)
Solute concentration refers to the amount of solute that is dissolved in a specified quantity of

solvent or solution (Hein and Arena, 2010). The quantity can be stated in terms of volume or

mass/molar amount (Pauling, 1988), while the concentration is usually stated in terms of

molarity (m), mass percent, and mole fraction (XA) as shown in equations (2), (3), and (4)

(Pauling, 1988).
moles of solute moles of solute
Molarity (𝑀) = Liters of solution (L) = Kilograms of solvent (kg) (2)
moles of substance, (𝐴)
Mole fraction (𝑋𝐴) = Total moles of solution (3)

31
Journal Pre-proof

moles of solute
Mass percentage of solute, (100%) = Mass of solution × 100 (4)

Table 13: Chemical composition of the sodium silicate solution (Malathy, 2009)
Composition Na2O SiO3 Water Specific Gravity pH
By mass, % 7.5 ± 8.5 25 ± 28 63.5 – 67.5 1.53 g/cc neutral

3 Clean production of GeoPC

GeoPC is usually produced by activating different alumino-silicate (Al–Si)-based waste

materials with a highly alkaline solution, such as alkaline earth metal silicate components, alkali

or alkaline earth metal hydroxide, fine and/or coarse aggregates, and water (Calderón-Moreno et

al., 2002). For example: inorganic polymeric ceramic formed from aluminum and silicon sources

that contain AlO4- and SiO4 tetrahedral units, under highly alkaline conditions. The ratio of SiO2

to M2O and to sodium hydroxide solids (NaOH) must be at least 0.8% and 97%–98% of the total

volume of GeoPC materials, where the purity of the NaOH in pellet form was 98%, respectively

(Calderón-Moreno et al., 2002; D Hardjito et al., 2005; Johnson, 2007). The concrete is initially

cast into a mold, after which it consolidates (Fig. 13). The ratio of alkaline liquid mass to FA

mass typically ranges between 0.30% and 0.45% (D Hardjito et al., 2005).

3.1 Composition of GeoPC

Geopolymer cement is used as a substitute to typical OPC (Komnitsas and Zaharaki, 2007). The

manufacturing of geopolymer cement needs an Al–Si precursor material (e.g., FA), a user-

friendly alkaline reagent (Na or K soluble silicates with a molar ratio (MR) of SiO2:M2O≥1.65,

where M can either be Na or K (Fig. 13) (Adam, 2009; A. M. M. A. Bakri et al., 2011). Room

temperature hardening can be easily achieved by adding a source of calcium cations, such as

blast furnace slags. Commonly, the first alkali-activated FA-based GeoPC needs heat hardening

at 60-80°C; furthermore, it cannot be formed separately but it can improve part of the FA-based

32
Journal Pre-proof

concrete. NaOH (user-hostile) + FA: FA particles entrenched in an alumino-silicate gel with

Si:Al= 1 to 2, zeolitic sort (chabazite-Na and sodalite).

Fig. 13: Constituents used in the production of GeoPC (Hassan et al., 2019)

The second slag/FA-based geopolymer cement requires head hardening at ambient room

temperature. The produced silicate solution + blast furnace slag + FA: FA particles are

entrenched in a geopolymeric matrix with Si:Al= 2, (Ca,K)-poly(sialate-siloxo) [37, 152].

Geopolymer cements (FA, GGBS) can be cured faster than PC, and some mixes can even reach

their maximum strength within 24 hours (A. M. M. A. Bakri et al., 2011). The obtained

33
Journal Pre-proof

compressive strength was between 60 and 70 MPa at 28 days (for high early strength production,

20 MPa after 4 hours and 25 MPa after 25 hours). However, these constituents must be set

slowly enough in order for them to be mixed with fine and/or coarse aggregates at a batch plant

to create GeoPC concrete as shown in Fig. 13.

3.2 Curing regimes of GeoPC

The curing of freshly prepared GeoPC is the most important part of the entire geopolymerization

process because of its key role in maximizing the quality of concrete (Chithra and Dhinakaran,

2014). Proper curing can also positively influence the final properties of GeoPC (Patil et al.,

2014). GeoPC is often cured at elevated temperatures in three ways, namely, steam, ambient, and

oven curing regimes. The result of each curing regime on the compressive strength of the GeoPC

is presented in Table 14.

3.2.1 Ambient curing

Ambient curing is performed after casting the specimens and letting them rest for a single day at

room temperature of 20±3 °C (Kumaravel, 2014; Nath and Sarker, 2012). Rest period refers to

the time from the completion of specimen casting to the initiation of curing at raised

temperatures (Kumaravel, 2014). In ambient curing, the compressive strength of GeoPC rises

from age 7 to 28 days (Zhuang et al., 2016b). GeoPC can be cured without using elevated heat

and could be applied to other areas beyond precast members (Nath and Sarker, 2012).

Furthermore, the addition of slag in the FA-based GeoPC mixture enhances the compressive

strength and reduces the setting time (Adam, 2009; Kumaravel, 2014; Phoo-ngernkham et al.,

2014). Also, the addition of slag by around 30% of the overall binder results in compressive

strength of approximately 55 MPa at 28 days while the setting time condenses quickly with

34
Journal Pre-proof

higher amount of slag in the mixture and the slump of fresh concrete reduces a bit when the slag

content increases (Deb et al., 2016; Nath and Sarker, 2012).

Table 14: Influence of each curing regime on the compressive strength of GeoPC
Types of Compressive Ref.
Curing regime Major findings
GeoPC strength, MPa
Cured at room temperature after Optimum strength
RHA (Ukwattage et al.,
casting for 2–12 achieved at 35 days of
based 2013)
14, 28, 35, 42, and 49 days curing
- Steam curing (5 h humidity
cabinet curing followed by
100°C steam curing for 8 h) Steam-cured specimens
GGBS- - Autoclaved curing (24 h exhibited higher (Zhang et al.,
15–90
based humidity cabinet curing strength than autoclaved 2016)
followed by 210°C, 2.0 MPa specimens
autoclaved curing for 8 h), at
28days.
POFA- 7 days 60–120°C oven curing after 90°C oven curing (Sivaraja et al.,
4–34
based casting optimal 2010)
24 h pre-curing period after
casting followed 80°C oven curing
49–60 (Ye et al., 2016)
by 36 h 50–90°C oven curing, at optimal
28days.
24 h 65°C curing; 5 min
5 min microwave curing (Dharmendra S.
microwave curing +
20–42.5 + 6 h 65°C curing Ravat and Dave,
3/6/12 h 65°C cures; ambient
optimal 2017)
temperature curing, at 14, 28 days.
FA based
60°C oven curing
25°C, 40°C, and 60°C curing for
optimal (applicable for 7 (Nath and Sarker,
24 h after 1 h of pre-curing 22–53
and 28 days strength 2012)
period, 14, 28, 49 days.
development)
10–12 h room temperature curing Sealed-condition curing
upon casting, followed by saline- optimal, followed by (Soroka et al.,
49–91
water, normal-water, and sealed- saline-water and normal- 1978)
condition curing, at 28, 58 days. water curing

3.2.2 Steam curing

This form of curing is done in autoclaves at temperatures in the 160 °C to 190 °C range and

pressures of 0.55 to 1.70 MPa. The condition changes the interaction of the hydration making a

concrete that has better sulfate resistance, no efflorescence, less shrinkage and creep, and lower

moisture content after curing (Chennur Jithendra, 2017; Komnitsas and Zaharaki, 2007). Steam

curing is performed after dismantling the specimens from their steel molds (after 24 hours) and

35
Journal Pre-proof

immediately storing them in a vacuum bagging film to generate steam at temperatures of 40 °C

to 100 °C (Kumaravel, 2014; Soroka et al., 1978). By reducing drying shrinkage (DS) and creep,

this curing regime is particularly useful in cold weather or when trying to achieve early strength

gain (D Hardjito et al., 2005; He et al., 2013). Steam curing can be performed under low

atmospheric pressure and high pressure in autoclaves (Aldea et al., 2000). The specimens are

placed inside the vacuum bagging film for 7 to 28 days (Adam, 2009; Shaikh, 2016), and the

temperature is utilized as a compromise between the ultimate strength and the rate of strength

gain (Hemalatha and Ramaswamy, 2017). In GeoPC, creep and dry shrinkage could be reduced

by adding the least volume of water possible in the concrete mix, increasing the relative humidity

of air, minimizing the cement paste volume, reducing the water/binder ratio, and adding large

coarse aggregates and a little bit of steel fibers (Duan et al., 2016; Kirupa and Sakthieswaran,

2015; Ridtirud et al., 2011; Yan et al., 2016).

3.2.3 Oven curing

In oven curing, the temperature in a hot air oven is maintained between 40 °C and 120 °C for 24

hours (Vijai et al., 2010). Afterwards, the oven is turned off to permit the cubes to cool down at

room temperature; next, samples are removed from the oven and tested for compressive strength.

The oven cured samples can improve the compressive strength of FA-based GeoPC under

elevated temperatures of up to 800 °C and oven curing regime at 80 °C by 2% to 7% along with

age (Kong and Sanjayan, 2010). Therefore, the strength gains in ambient curing are considerably

larger than those in hot curing (Ganesh et al., 2016; Nath and Sarker, 2012; Vijai et al., 2010).

Reportedly, the rate of strength gain was found lower at 60°C in comparison to strength at

120°C; thereby, a compressive strength higher than 60 MPa can be attained by FA-based GeoPC

in only 24 hours of curing (Chithra and Dhinakaran, 2014; Kumaravel, 2014; Patil et al., 2014).

36
Journal Pre-proof

In addition, the curing regimes perform a vigorous role in the development of strength and the

micro-structural system of GeoPCs, meaning that when samples are subjected to elevated

temperature, they can attain higher strength.

4 Fresh properties

Fresh GeoPC practical requirements differ from those of vibrated fresh concrete. A concrete

mixture is only categorized as GeoPC if its three key features, namely, resistance to segregation,

passing ability and filling ability are all satisfied according to European guidelines [165]. The

workability, stability, and flowability concerning the fresh properties of various GeoPC mixes

are commonly evaluated using slump flow, T50cm slump flow, V-funnel, L-box, and J-Ring test

procedures (Table 15) (Chennur Jithendra, 2017; Druta, 2003; Subang Jaya et al., 2013; Ushaa et

al., 2015). The fresh properties of GeoPC, including its stability, compatibility, setting time,

workability, and deformability, are described in the following subsections.

Table 15: Limited values of fresh GeoPC tests according to the EFNARC guideline (Chennur
Jithendra, 2017; EFNARC, 2002)
Typical tests Units Minimum Value Maximum Value Refs.
Slump flow mm 650 800 (Aharon-Shalom
T50cm slump flow Second 2 5 and Heller,
J-ring mm 0 10 1982; Chennur
V-funnel Second 8 12 Jithendra, 2017;
L-box H2/H1 0.8 1.0 Subang Jaya et
U-box (H2-H1) mm 0 `30 al., 2013)

4.1 Stability

Stability is the ability of a material to keep its original formation and structure when facing

various impacts, including environmental impacts (Gaochuang et al., 2016). A viscosity-

modifying admixture (welan gum) ensures the sufficient stability of concrete cast in deep

structural members (Khayat et al., 1997). GeoPC mainly consists of covalent bonds and offers

several advantages, such as low density, excellent volume stability, and ability to avoid

37
Journal Pre-proof

deformations (Su et al., 2016). Limestone powder is frequently used as an inert mineral additive

to reduce the content of energy-intensive reactive binders and to enhance the stability of fresh

concrete mixtures (Triantafillou, 2016). Using a large quantity of fine fillers and/or additives to

increase viscosity can also maintain the mixture design of GeoPC, reduce bleeding, and prevent

the separation of coarse aggregates (Ushaa et al., 2015). Hawes et al. (Hawes et al., 1992, 1990)

found that adding pozzolans, such as silica fume and FA, can increase the stability of concrete

that incorporates phase change materials. Memon et al. (Memon et al., 2013) reported that

GeoPC becomes unstable and weak when the amount of silica fume exceeds that of FA by over

10%. Using super-plasticizers (SPs) can also influence FA-based geopolymer pastes at varying

degrees because of the instability of these commercial materials in highly basic media, such as

NaOH+Na2SiO3 (Memon et al., 2012; Palacios and Puertas, 2004). In GeoPC, it is reported that

when the use of hybrid fibers, 80% GGBS and 20% RM, a 30% upsurge in flexural strength

takes place as compared to normal concrete (Adam, 2009). Samples achieved their properties at

ambient curing temperature of 40-50 oC. Hence, GeoPC strength increased by about 23% when

used 17% GGBS parallel with 5% Na2SO4 and 5% H2SO4 whereas normal concrete suffered

losses. The presence of Al3+ in controlled Si/Al–Ca/Al or Si/Al–Na/Al ratios can positively

influence the chemical stability and durability of the GeoPC matrix (Kamseu et al., 2016), and

reduce the amount of sulfur trioxide (SO3) by less than 1%; thereby, ensuring high volume

stability (Gunasekara et al., 2016). Commonly, the stability of GeoPC can be enriched by

decreasing its paste volume and cement content. Furthermore, aggregate typically suffers from

shrinkage; thereby, the tendency of stability can be upgraded by destroying the contraction of

aggregate.

38
Journal Pre-proof

4.2 Compatibility

GeoPC is well-known to be compatible and robust when it comes to interaction between the

various components in the mix design and other chemical admixtures (Anuradha et al., 2014;

Kavitha et al., 2016; Sashidhar et al., 2016). Consequently, where there is no cooperation

between various components in the matrix, the compatibility of GeoPC mortar would be

condensed (Chennur Jithendra, 2017; Druta, 2003; EFNARC, 2002; Subang Jaya et al., 2013).

Hence, segregation may frequently occur when contact issues arise between the plasticizers and

surfactant due to the incompatibility of the admixtures in the mix design (Kamseu et al., 2016;

Memon et al., 2013, 2012). GeoPC is synthesized from low-calcium FA and triggered by a

combination of sodium silicate and sodium hydroxide solutions. Also, incorporating SPs

improve the self-compatibility and influence the hydration of GeoPC (Ganesh et al., 2016; Han

et al., 2014). In GeoPC, the self-compatibility property is essentially obstructed by the features of

materials (e. g. super plasticizer) and also the proportions of the mix; henceforth, it becomes

important to develop a procedure to improve mix design of self-compacting GeoPC (Huseien et

al., 2015). A self-compacting concrete achieves consistency and self-compatibility under its own

weight without the need for any external compaction and can be influenced by the characteristics

of materials and the mix proportions of various byproduct materials (e.g., RHA, FA, and SFA);

therefore, a procedure for ensuring a suitable mix design for GeoPC must be proposed

(Moghadam and Khoshbin, 2012; Prabhu et al., 2016). In the limited oxygen index test, the

alkaline-treated coconut tree leaf sheath fiber with phenol formaldehyde resin shows high

compatibility in composite geopolymer (Bharath and Basavarajappa, 2014). Using FA as a

substitute for 30% of GGBS can also increase the compressive strength by 60% at 28 days and

compatibility of GeoPC in the fresh state by 12% (Ganeshan and Venkataraman, 2017).

39
Journal Pre-proof

Lightweight aggregate concretes are usually linear to levels approaching 90% of the failure

strength, thereby showing the relative compatibility of the constituents and reducing the

formation of micro-cracking (Chandra and Berntsson., 2002). Al-Mulla (Al-Mulla, 2010)

reported that a 10% RHA absorbs a large amount of water from the mixture; thereby, reducing

the concrete strength due to inadequate compatibility. Meanwhile, those binary cements that

contain 40% to 50% FA, nearly 10% silica fume, and 50% to 70% GGBS show high

compatibility with GeoPC in the fresh state, increased sulfate resistance in the hardened state,

and relatively slow strength development under normal or low temperatures (Gjorv and Sakai.,

1999). Thus, it can be concluded that the performance of the self-compatibility of GeoPC relies

on the ratio of fine and coarse aggregates (50% by volume), super-plasticizer dosage (≤ 2%) and

the w/c ratio (0.9-1.0) (Huseien et al., 2015).

4.3 Setting time

Geopolymer mortar has an initial setting time of 35 min and a final setting time of 600 min at

curing temperatures of 20 °C to 80 °C (Vijai et al., 2010). Low-calcium FA has a much higher

setting time than high-calcium FA. Those paste samples with some formulation of alkali-

activated FA have a setting time of less than 5 min to 7 min, which can reach 20 min to 40 min

in the best case. Meanwhile, those slags and other materials activated with sodium hydroxide

require less than 3 min to 4 min and less than 15 min to activate sodium silicate, respectively.

Reportedly, the setting time of GeoPC mixtures is having the same shares of alkaline activator

and binder of the conforming mortar mixtures with the fine aggregate misplaced (Nath and

Sarker, 2015). It is known that the sodium hydroxide reaction in solution is exothermic, and it is

suggested that the mix design of Na-Si and NaOH should be completed one day before mixing

with the calcio-aluminosilicate constituents (Subang Jaya et al., 2013). This will guarantee

40
Journal Pre-proof

equilibration and ensures that no unrestricted heat during mixing that will affect setting of the

GGBS-based GeoPC paste (Castel and Foster, 2015; Dimas et al., 2009; Jain and Pal, 1998;

Suresh and Nagaraju, 2015). It is reported that when NaOH is used as an activator, the concrete

mechanical properties rise with the reduction in water/binder ratio. Another research on FA-

based GeoPC revealed that the increase in Na2SiO3/NaOH ratio to 2.5 contributed to a

considerable decrease in mortar setting time (Ali and Zurisman, 2015). To examine the

microstructure of OPC blended with FA-based geopolymers, two paste mixes that contained

OPC as a substitute for 10% and 50% of the total binder were prepared (Adam, 2009; D Hardjito

et al., 2005; Hardjito et al., 2008). An alkaline solution with a 2.5 (Na2SiO3/NaOH) ratio and a

2.0 solid/liquid ratio were adopted as a substitute for 40% of the overall binder. It was observed

that the initial setting time was doubled as a result of increase in alkaline liquid content from

35% to 40%. The samples were cured at room temperature (20 °C to 23 °C) to which a slow

setting was observed at this temperature using a Vicat apparatus (Nath and Sarker, 2012). In

addition, using GGBS and silica fume improve geopolymerization by reducing the setting time,

while using alternate materials, such as RHA, class C FA, metakaolin, and RM, also leads to

positive results (Ganeshan and Venkataraman, 2017). Generally, the setting time of GeoPC can

be affected by numerous factors such as chemical and physical properties of the binder itself,

mix design composition, mixing process, molarity of NaOH, Na2SiO3/NaOH ratio and ecological

conditions.

4.4 Workability

The amount of extra water in GeoPC (a water content of 12% and SP dosage of 6% by mass) is

an important gradient required for regulating strength and workability (Ganeshan and

Venkataraman, 2017; Ushaa et al., 2015). Nath and Sarker (Nath and Sarker, 2015) stated that

41
Journal Pre-proof

increasing the alkaline liquid content can lower the strength and improve concrete workability

because of the high liquid–solid ratio of the mixture with the highest liquid content. Therefore,

using a mixture with 5% OPC content can achieve a reasonably high early age strength, setting

time, and workability for ambient-cured FA-based GeoPC (Nath and Sarker, 2012). Another

research on FA-based GeoPC revealed that the increase on Na2SiO3/NaOH ratio to 2.5 resulted

in a substantial reduction in the workability of the mortar (Ali and Zurisman, 2015). It is also

reported that adding naphthalene-sulfonate-based SP, the use of total 4% of SP per FA by mass,

led to an enhancement of the workability of GeoPC in the fresh state; however, the compressive

strength of GeoPC in the hardened state slightly declined when the SP dosage exceeded 2% (D

Hardjito et al., 2005). Nevertheless, the workability of GeoPC in the fresh state is remarkably

enhanced by adding silica fume up to 10% (Memon et al., 2013). Reducing the particle size can

also improve the workability of the mixture (Leong et al., 2016), while the addition of fibers can

reduce the workability (Kovler and Roussel, 2011). The addition of naphthalene-based SP in

GeoPC can influence both the workability and final setting time depending on the percentage

added between 1.5 to 3% of FA by mass (Memon et al., 2012; Nematollahi and Sanjayan, 2014).

It is also observed that the workability of FA-based GeoPC with SFA improved when SP was

limited to1.5%, which led to enrichment of the compressive strength and showed a slight

degradation when the SP value was 2%. Sakulich (Sakulich, 2011) found that adding slag

enhanced the workability of the matrix because slag needs a lesser amount of liquid than

metakaolin for particle wetting, which helps in reducing the porosity and water permeability. In

this case, those GeoPC mortars blended with flash metakaolin have better workability than those

blended with rotary-kiln metakaolin (San Nicolas et al., 2013). Among several mineral

admixtures, the blast furnace slag series have better workability than the silica fume series

42
Journal Pre-proof

(Ushaa et al., 2015), while the fine aggregates fill the voids and increase the workability of

concrete (Sashidhar et al., 2016). Behzad and Sanjayan (Nematollahi and Sanjayan, 2014) found

that using different commercial SPs (1.0%) does not enhance the workability of the activated FA

pastes in GeoPC. Furthermore, the proportion of GGBS and silica fume in GeoPC rises to more

than 30% and 15%, respectively, and does not meet the workability requirements of this concrete

(Anuradha et al., 2014).

4.5 Deformability

Knowledge on the deformability of concrete is highly imperative in the calculation of structural

deflections, computation of strains and stresses, and in the development of constitutive models

for simulations. The concrete digital dilatometer (CDD) (Fig. 14) is one of the tools used to

measure the linear deformation of GeoPC in both fresh and hardened states (Esping and

Löfgren., 2005). Meanwhile, slump flow-rate is a fast and a simple procedure that is commonly

used in laboratories and construction sites, and it applies horizontal free flow measurements of

the concrete in the fresh state (Memon et al., 2013). This method can favorably assess the

deformability or flowability of concrete in the fresh state and can balance deformability with

stability (Sashidhar et al., 2016). In addition, it is reported that the hardness of GeoPC is almost

two times higher than that for normal concrete, and it could show higher brittleness and less

deformability (Kabir et al., 2015; Saraya, 2014). Because the chemical admixtures of GeoPC can

improve the deformability and viscosity, this causes the mixtures to have high filling capacities

ranging from about 60% to 70%; thereby, indicating exceptional deformability without blockage

within obstacles that are closely spaced (Aggarwal et al., 2008; Khayat and Guizani, 1997) . A

low water content necessitates a relatively high dosage of high-range water reducers to gain the

desired deformability (slump flow, 660 mm to 690 mm) especially when lower binder contents

43
Journal Pre-proof

are available because of high paste viscosity and high inter-particle friction (Sashidhar et al.,

2016; Ushaa et al., 2015). Previous studies were performed to measure the behavior of GeoPC,

including the influences of chemical admixtures, C-S-H phase and curing conditions. It was

found that the addition of FA/GGBS-based GeoPC pastes, alumino-silicate gel (N-A-S-H) and

C-S-H; activated principally by NaOH at a low temperature of 27 oC, thereby, the concrete paste

is led by the N-A-S-H and C-S-H, depending on the alkalinity volume of activators used (Bakri

et al., 2012; Görhan and Kürklü, 2014; Kamseu et al., 2016; Sashidhar et al., 2016; Singh et al.,

2015). Rickard et al. (Rickard et al., 2016) found that expanded clay aggregates (quartz content

<30%) have higher deformability than quartz aggregates. The presence of polypropylene fiber

(0.1%) reduces the deformability of GeoPC by increasing the surface area that must be lubricated

by cement paste and water (Muthupriya et al., 2014). Given that a GeoPC with a high w/binder-

ratio (0.67) has a large crack area, the w/binder-ratio in the region must be at least 0.55 (Esping

and Löfgren., 2005). However, it can be concluded that decent understanding on the

deformability of GeoPC is an essential key towards the estimation of deflection curves, and the

stress-strain relationship in order to develop the solid rules for finite element modeling.

44
Journal Pre-proof

Fig. 14: Test arrangement of the CDD for linear deformation measurement (Esping and
Löfgren., 2005)
5 Mechanical properties

After setting, it is necessary to ensure that the concrete is sufficiently hard to resist the applied

service and structural loads. However, given that GeoPC is produced of high-quality materials

and is suitably proportioned, mixed, handled, placed, and finished, this concrete is among the

most durable and strongest construction materials available in the market. The mechanical

properties of GeoPC, including its compressive, splitting tensile and flexural strengths, modulus

of elasticity (MoE), stress–strain behavior, and rate of strength development, are reviewed in the

following subsections.

5.1 Compressive strength

GeoPC’s compressive strength (ASTM C39) is affected by wet-mixing time, curing time, curing

temperature, particles size (Chang and Shih, 2000; Lakshmi and Nagan, 2011) and addition of

typical additives (Ca(OH)2, Al(OH)3, SF (Ye et al., 2016), nano-SiO2 (Boonserm et al., 2012),

nano-Al2O3 (Nath and Kumar, 2013), vinyl (Nematollahi and Sanjayan, 2014) and copolymer

and polyacrylate copolymer based SP (Puertas et al., 2003). The impact of these additives on the

compressive strength of GeoPC is summarized in Table 16. FA-based GeoPCs exhibit a steady

reduction in their original compressive strength at elevated temperatures of up to 400 °C

regardless of their molarities and coarse aggregate sizes (ASTME119, 2012; Chu et al., 2016;

Görhan and Kürklü, 2014; Kong and Sanjayan, 2010). The compressive strength of GeoPC also

reduces considerably when the amount of extra water exceeds 12% of the FA mass (Memon et

al., 2012). Meanwhile, calcination slightly changes the mineralogical composition of this

concrete and reduces its compressive strength by almost 30% (Temuujin et al., 2011). Replacing

40% of cement with GGBS can cause greater improvements in compressive strength compared

45
Journal Pre-proof

with a 20% or 60% replacement (Chithra and Dhinakaran, 2014; Nath and Sarker, 2012).

Numerous scholars have examined the mechanical properties of GeoPC that contain 50% and

100% RCA and deduced that the compressive strength of this concrete rises by approximately

10% from 7 to 28 days (Brake et al., 2016; Nuaklong et al., 2016; Shi et al., 2012; Talakokula et

al., 2016). Such compressive strength reaches its peak when a mixture of water-glass and NaOH

is adopted as the activator but decreases when the alkaline solution content increases from 35%

to 45% of the total binder (Nath and Sarker, 2015). The sodium alumina silicate hydrate (N–A–

S–H) gel for 100% FA can also improve the compressive strength of GeoPC (Soutsos et al.,

2016). To achieve a microstructural phase of NASH gel, or KASH when using potassium, it is

stated that the strength development of GeoPC mixtures based on sodium hydroxide and calcined

kaolin have shown lesser reactivity than the mixtures in which this ast component was

substituted by potassium hydroxide (Nuaklong et al., 2016; Shi et al., 2012). Besides, the

findings display that there is no direct linear bond between the strength and the calcium

hydroxide content. Vaidya and Allouche (Vaidya and Allouche, 2011) found that the addition of

fiber can meaningfully enhance both the ultimate flexural capacity and ductility of FA-based

geopolymers, particularly at the early ages, without adversely influencing their ultimate

compressive strength. Such influence can be maintained by replacing FA with 30% palm oil FA

(POFA), 30% GGBS, and 10% SFA (Anuradha et al., 2014; Ganeshan and Venkataraman, 2017;

KThu and Murthy, 2015; Mo et al., 2015a; Prabhu et al., 2016). Bakrilso (Bakri et al., 2012) and

Sanjayan (Nematollahi and Sanjayan, 2014) deduced that the compressive strength of concrete

reduces by 29% when the optimum Na2SiO3/NaOH ratio is 2.5. The correlation between the

compressive strength and the SiO2/R2O ratio indicates that a rise in alkali content or a reduction

in silicate content can increase the compressive strength of geopolymers by forming Al–Si

46
Journal Pre-proof

network structures (Singh et al., 2015; van Jaarsveld and Deventer, 1999). Applying a silicate

solution to GeoPC can increase its density from 79% to 93% and its compressive strength by

35% at room temperature (He et al., 2010; Sakulich, 2011). However, the compressive strength

of this GeoPC reduces with the addition of fiber (e.g., glass, carbon, polyvinyl chloride, and

polyvinyl alcohol) (Al-Majidi et al., 2017; Karbhari, 2013; Kovler and Roussel, 2011; Li and Xu,

2009; Rickard et al., 2014). Furthermore, previous researchers indicated that the compressive

strength of RCA-based GeoPC that uses geopolymer from wastepaper sludge ash instead of that

from FA and slag rose by roughly 10% from 7 to 28 days. Moreover, the high molarity of

sodium hydroxide indicated a higher compressive strength in GeoPC than that in conventional

concrete (Anuar et al., 2011; Nuaklong et al., 2016; Posi et al., 2013; Sata et al., 2013; Shi et al.,

2012). Also, the addition of slag by around 30% of the total binder attained a compressive

strength of approximately 55 MPa at 28 days. The setting time condensed rapidly with greater

volume of slag in the mixture, and the slump of fresh concrete reduced a bit when the slag

content increased (Deb et al., 2016; Nath and Sarker, 2012). Using FA as a substitute for 30% of

GGBS can also increase the compressive strength by up to 60% at 28 days and increase the

compatibility with GeoPC in the fresh state by 12% (Ganeshan and Venkataraman, 2017).

Moreover, it is reported that the compressive strength of FA-based GeoPC only shows a 12% to

40% reduction when exposed to AE, while OPC concrete shows about 40% to 65% reduction in

compressive strength under the same conditions and time (A. Castel, 2016; Azreen et al., 2018).

GeoPC has potential to produce ultra-high compressive strengths with better durability

performance, leading it to be used in the fabrication of concrete applications that suffer from

aggressive environments. According to Neville (Neville, 1995), the relationship between the

splitting tensile and compressive strengths of OPC concrete can be given as

47
Journal Pre-proof

fct = 0.3 (fcu)2/3 fct - principal tensile strength of concrete, N/mm2; and (5)
fcu - compressive strength of concrete, N/mm2.
𝑛
1 K - Empirical constants, n = Strength to gel-space ratio (6)
S=K [ 𝑎 ]
(1 + ) + (( )
𝑤
𝑐 𝑐

S =PO (1 - P)n So - The strength at zero porosity (7)


n - Balshin’s constant
𝑃𝑐𝑟
S =Ks ln [ 𝑃 ] Pcr - Critical porosity at zero strength (8)
Ks - “Schiller’s constant
S = Kgn K - The intrinsic gel strength (9)
g - The gel-space ratio (Power’s gel-space ratio)

Table 16: Influence of additives on the compressive strength of GeoPC


Types of Compressive
Types of additives Major findings Ref.
GeoPC strength, MPa
15–25% Ca(OH)2, 5–
POFA- 20% Ca(OH)2, 5% SF and 10%
10% Al(OH)3, 2.5–7.5% 15.67–44.74 (Shi et al., 2015)
based Al(OH)3 optimal
SF
ASTM Class C FA (0, (Ryu et al., 2013)
RHA- 50% FA and 5% GGBS
25, 50, 75, 100%), 5–55
based replacement level optimal
GGBS (0, 5, 10, 15%)
0–50% GGBS 50% GGBS replacement optimal
(Görhan and
replacement by weight 8.5–93.4 and FA-GCS yielded higher
Kürklü, 2014)
of binder strength than FA-GGBS
1–3% of nano-SiO2 and (Pauling, 1988)
2% nano-SiO2 and 1%
nano-Al2O3 by binder 20.2–56.4
nano-Al2O3 optimal
weight
1% addition by mass of PC-based SP showed highest (Adam, 2009)
binder of N, M, and PC 47–81.3 plasticizing effect and least strength
FA- based SP reduction
based 0.5–1.5% addition of (Sata et al., 2012)
vinyl Insignificant changes in
copolymer and 30–35 strength and workability
polyacrylate copolymer with addition of both SPs
based SP
Addition of 2.5–5% of (Chindaprasirt et
Al-rich waste with 2.5% and
Al-rich waste calcined at al., 2014)
27.4–34.2 1000°C calcined temperature
400, 600, 800, and
optimal
1000°C
RM 10-40% RM by wt of 25% was found the optimum (Poudenx, 2008)
50.91-27.74
based cement percentage of replacement.
The value of MoE for 90% of (Fernandez-
GGBS GGBS was replaced up
Up to 30 normal concrete using the same Jimenez et al.,
based, to 40%
type of aggregate. 2006)
The strength was increased by 84%, (Anuradha et al.,
SiO3/OH = 0.5, AL/SF =
38% and 15% at 7, 28 and 56 days, 2014)
SF based 0.25, up to 56-60% of 10.31-37.5
respectively, for the whole AL/SF
SF by wt of cement
ratios.

48
Journal Pre-proof

5.2 Splitting tensile strength

The splitting tensile strength (STS, ASTM C 496) is a basic and an important property of

concrete which is weak in tension due to its brittle nature and is not typically designed to

withstand direct tension (Druta, 2003). The STS of self-compacting concrete is assumed to be

almost 30% more than that of normal concrete (Anuradha et al., 2014; Chennur Jithendra, 2017;

Druta, 2003; Kavitha et al., 2016; Neville, 1995; Subang Jaya et al., 2013; Ushaa et al., 2015).

Using slag as a partial substitute for FA can improve the STS of GeoPC, while incorporating

10% and 30% RCA (KThu and Murthy, 2015; Mo et al., 2015b; Nuaklong et al., 2016) or 10%

palm oil shell aggregate can reduce the STS (Liu et al., 2016; Mo et al., 2015a). An SFA content

of over 10% can also reduce the STS of GeoPC (Memon et al., 2013). Islam et al. (Al-Majidi et

al., 2017) found that POFA and GGBS can increase the STS of GeoPC by approximately 6%–

9% and 9%–11%, respectively. Siva Raja et al. (Sivaraja et al., 2010) examined the mechanical

properties of sisal-fiber-reinforced GeoPC at a three-month interval and found that sisal fiber

only enhances the STS of this concrete by 8.4%. Using FA and granite slurry can improve the

STS of concrete after 28 days by 7.8% and 40%, respectively (Ryu et al., 2013; Sreenivasulu, C.

and Jawahar, 2015). Maochieh and Huang (Chi and Huang, 2014) found that replacing fine

aggregates with circulating fluidized bed combustion ash can increase the STS of GeoPC by 5%

to 10%. Moreover, the use of 0.03 vol% steel, 10% synthetic, 2% sweet sorghum, 1%–5%

carbon, 0.5 vol% anon-metal, and 0.5 vol% polypropylene fibers can increase the STS of GeoPC

by 16%, 12.8%, 36%, 8.4%, 12%, and 14.4%, respectively (Bashar et al., 2016; Chu et al., 2016;

Karbhari, 2013; Mazaheripour et al., 2011; Rickard et al., 2014). A research on 90% FA/and

10% GGBS-GeoPC with 0.25% steel fibers under various curing conditions found that the STS

increases with an increase in the volume level of steel fibers (Chithra and Dhinakaran, 2014;

49
Journal Pre-proof

Nath and Sarker, 2012). Similar findings were reported when 1.5% steel fibers were added into

the mix design of slag-based GeoPC. The influence of these additives on the STS of GeoPC is

summarized in Table 17. The Australian standards for concrete structures (AS3600) (AS, 2009)

proposes the following equations for computing the characteristic principal STS (fct) of GeoPC:

ft = 0.20 (fc)0.70 For density between 1400 - 1800 kg/m3 (10)


ft = 0.23 (fc)0.67 fc= 28 days compressive strength N/mm2 (11)
fct = 0.4 𝑓𝑐𝑚 Where, P = load at failure (N); (12)
2P D = diameter of specimen (mm); and (13)
ft = πDL L = length of specimen (mm)
Further, the expression below is proposed for STS of fiber reinforced GeoPC with respect to the

volume fraction level of steel fibers (Vf) for oven-curing;

STS = fsc + 0.743 Vf (14)

Where, fsc = the STS of GeoPC composites at 28 days,


Vf = the volume fraction level of steel fibers.

5.3 Flexural strength

According to ASTM C 78, the flexural strength of geopolymers can be considerably improved

through the incorporation of synthetic fibers like polypropylene and PVA. The bridging effect in

the micro- and macro-cracking of the geopolymer matrix under flexure improves the interfacial

strength. However, an excessive addition of fibers can reduce the flexural strength (A. Castel,

2016; Brake et al., 2016; Breña et al., 2001; Sharafeddin et al., 2013). For instance, 2% addition

of sweet sorghum fiber improves the flexural capacity of ASTM class F FA-based geopolymer

specimens by about 40% (Chen et al., 2014). Also, 2% addition of nano-SiO2 and nano-Al2O3 by

weight improves the flexural strength significantly in a high-calcium FA-based geopolymer

specimens (Phoo-ngernkham et al., 2014). Meanwhile, adding 10% SF in an FA-based GeoPC

improves the flexural strength by 11.09% (Memon et al., 2013). The flexural strength of concrete

with an optimum addition of cotton fabric (8.3%) is nearly three times more than that of

50
Journal Pre-proof

unreinforced GeoPCs (Alomayri et al., 2014a). The flexural strength of GGBS-based GeoPCs

also increases by incorporating 1%–15% polymer resin (Zhang et al., 2010). The addition of 1%

resin greatly improves the flexural strength of GeoPCs by 41%. Prabhu et al. (Prabhu et al.,

2016) found that concrete’s flexural strength can increase by up to 25% when replacing cement

with 10% FA, 10% GGBS, and 5% SFA. The hybrid GeoPC with 40% to 80% PFA content has

a higher flexural strength than a control mixture with 0% PFA content (Dimas et al., 2009). The

impact of these additives on the STS and flexural strength of GeoPC is summarized in Table 17.

Also, research on FA-based GeoPC with addition of 10 and 8 NaOH molarities found that the

flexural strength increased by 3.5% as the concentration of NaOH molarities rose from 8 to 10

(Bakri et al., 2012; Somna et al., 2011). Another study found that the density of FA-based

GeoPC increased with a longer curing period but the degradation of higher temperature curing

condition caused a poly-reaction, leading to reduction in the density of samples at the same time

(Aldea et al., 2000; Atiş et al., 2005). But FA-based GeoPC with the addition of GGBC with

Phosphogypsum revealed an increase in flexural strength by 7.5% compared to normal concrete.

However, it was observed that the flexural strength of GeoPC is smaller when compared with

normal concrete excluding when a lesser amount of alkali activator solution as cement paste is

replaced, which can lead to attainment of higher flexural strength.

Table 17: Influence of additives on the STS and flexural strength of GeoPC
Types of Types of Splitting tensile Flexural
Major findings Ref.
GeoPC additives strength, MPa strength, MPa
41% enhancement (Mo et al.,
GGBS- 1–15% addition in flexural strength 2015b)
2.34- 3.66 4.8–8.6
based of polymer resin by 1%
resin addition
0–3% addition of (Pauling, 1988)
1% SiO2 and 2%
nano- SiO2 and 3.66–5.12
Al2O3 optimal
nano-Al2O3
FA-based 0–8.3% addition 4.14–4.67 8.3% and (Mazaheripour et
of horizontally horizontally al., 2011)
8–32
and vertically oriented cotton
oriented cotton fabric optimal

51
Journal Pre-proof

fabric
1, 2, and 3% Addition of 2% of (Mo et al.,
addition sweet 2015b)
2.2–3.4 M 3.2–5.6
of sweet sorghum sorghum fibers
fibers optimal
0–15% addition (Khayat et al.,
4.14–4.67 4.09–4.56 10% of SF optimal
of SFA 1997)

5.4 Modulus of elasticity

Modulus of elasticity (MoE) (ASTM C 469) is highly correlated with the compressive strength

of any concrete type (e.g. GeoPC), thereby a higher degree of geopolymerization can result in a

denser geopolymer matrix, which in turn leads to a higher MoE (Topark-Ngarm et al., 2014).

However, geopolymerization is known as the formation of polysialates that depend greatly on the

mineralogical, chemical and physical properties of the raw materials, amount of activator and

curing conditions (Ahmari and Zhang, 2012; Boonserm et al., 2012; Davidovits, 2015; Kumar

and Kumar, 2013). The MoE property does not rely totally on the dosage of the chemical

activator but is also governed by the amount of aggregates in GeoPC mixtures (Khandelwal et

al., 2013). Low MoE can reduce the degree of crack propagation initiated by the corrosion of

steel bars. A high fine aggregate–total aggregate ratio can produce a GeoPC with an equal or

higher MoE (A. Castel, 2016). Several studies have stated that high silicate content can raise the

MoE of GeoPC and reduce that of OPC concrete (Yusuf et al., 2015). Several studies of FA-

based GeoPC obtained MoE values for specimens between 23.0–30.8 GPa (D Hardjito et al.,

2005), 10.7–18.4 (Sumajouw et al., 2004) and 30.3–34.5 GPa (Hardjito et al., 2008). Also, it was

reported that the MoE of pulverized fuel ash (PFA) mortars was lower than OPC mortar due to

the presence of alkali activated PFA mortar, but later at a long-term, the MoE of PFA mortars

increased to about 5–20% higher than OPC mixes (Dimas et al., 2009; Palacios and Puertas,

2004; Puertas et al., 2003). Ngarm et al. (Topark-Ngarm et al., 2014) stated that high-calcium

52
Journal Pre-proof

FA-based GeoPC displays equal or greater MoE with a low sodium silicate–NaOH ratio, which

corresponds to a high amount of Na2O. The incorporation of various fibre types, polymer resin,

SPs, and nano-materials can significantly improve geopolymers mechanical properties, including

their flexural strength, STS, and MoE (Alomayri et al., 2014b; Chen et al., 2014; Nematollahi

and Sanjayan, 2014; Phoo-ngernkham et al., 2014). Furthermore, adding 2% nano-Al2O3 and

nano-SiO2 by weight of binder can improve the MoE by about 30% of high-calcium FA-based

geopolymer samples cured at ambient temperature. After 90 days of curing, the E-value of these

samples can reach to as high as 17.65 GPa, which is comparable to that of OPC concrete (Phoo-

ngernkham et al., 2014). Moreover, further research on FA-based GeoPC found that the MoE

was 15-28 % lesser in comparison to normal concrete due to the addition of a low volume of

silicate content and sodium hydroxide solution in GeoPC (Fernandez-Jimenez et al., 2006). Also,

the combination of FA-based- and GGBS based GeoPC at ambient temperatures resulted in MoE

values between 10 GPa and 21 GPa, which is parallel with the compressive strength of 30 MPa.

However, the value of MoE for GeoPC is almost 90% of normal concrete using the same type of

aggregate. MoE can be computed using following equations;

E = 33 W1.5 (f'c)0.5 It was used Pauw’s equation (15)


𝐸𝑆𝐶𝐺𝐶 = ― 11400 + 4712 𝑓𝑐𝑢 fcu = compressive strength of (16)
FA-based GeoPC at 28 days.
E =5.31 × W - 853 Density is ranged between 200 - 800 kg/m3 (17)
Ec = 9.10 (fc)0.33 fc = compressive strength of concrete
Ec = 1.70 × 10-6 Ρ2 (fc)0.33 Ρ = plastic density (kg/m3) (18)
E =0.99 (f'c)0.67 It used when FA utilized as fine aggregate (19)
E =0.42 (f'c)1.18 It used when Sand utilized as fine aggregate (20)
5.5 Stress–strain behavior

The stress–strain behavior of GeoPC materials (ASTM C 469) must be investigated to

completely categorize their performance under field implementation and design (Noushini et al.,

2016). Investigations have shown that the stress–strain behavior depends on the type of concrete

53
Journal Pre-proof

and confinement (Ganesan et al., 2014; Haider et al., 2014). Hardjito et al. (D. Hardjito et al.,

2005) examined the stress–strain behavior of three sodium silicate FA-based geopolymer

matrixes containing 408 kg/m3 FA at 28-days compressive strength of 41 MPa to 64 MPa. The

stress–strain behavior of heat-cured specimen at 48 h and 50 °C were observed. The results

showed that FA and slag-based GeoPCs with high FA/GGBS content between 570 to 620 kg/m3

and Na–silicate solution with MoE of 0.75<Ms<1.5 indicated more brittle fracture compared

with PC concrete [236]. More investigations on the stress–strain behavior of confined GeoPCs

pastes have witnessed a tremendous increase over the years. Haider et al. (Haider et al., 2014)

carried out empirical investigations and observed the stress–strain behavior of sodium silicate

FA-based geopolymer paste under constant levels of confinement. It was determined that

geopolymer paste indicated lower deformation in the axial direction compared with OPC

concrete under similar confinement (A. Bakri et al., 2011; Malathy, 2009; Nath and Sarker,

2015; Noushini et al., 2016; Shi et al., 2011; Venkatanarayanan and Rangaraju, 2014). Ganesan

et al. (Ganesan et al., 2014) carried out investigative comparisons between OPC concrete and

confined sodium silicate FA-based GeoPC and observed their stress–strain behavior. Initially,

GeoPC samples were cured at ambient temperature over 24 h before heating at 60 °C in an oven

for another 24 h. It was observed that the stress–strain model proposed in the literature for

confined OPC can serve for GeoPC through curve fitting and adjustment of the curve factor

(Mander et al., 1988). Furthermore, it has also been determined that the stress-strain behavior of

GGBS/RHA-based GeoPC under compression is similar to that of normal concrete (Mo et al.,

2015a; Nematollahi et al., 2017; Noushini et al., 2016). Also, adding 0.05% Polypropylene fibres

was determined to increase the stress-strain by 60% at peak stress of GeoPC (Chu et al., 2016;

Mazaheripour et al., 2011; Rickard et al., 2014). Another study of RM-based GeoPC with

54
Journal Pre-proof

sodium silicate solution found that the stress-strain curve increased at the peak, indicating a

strong and a more ductile behavior of the GeoPC matrix (Ganesan et al., 2014; Haider et al.,

2014; He et al., 2013; PARAMGURU et al., 2004; Thomas and Peethamparan, 2015). Also, the

increase in RCA content was reported to slightly improve the ultimate axial stress-strain

behavior of the unconfined GeoPC samples because of the increase in the strain softening

(Ganesan et al., 2014; Haider et al., 2014; KThu and Murthy, 2015; Shaikh, 2016; Shaikh et al.,

2015; Thomas and Peethamparan, 2015). Collins et al. (Collins et al., 1993) suggested that the

stress–strain behavior of concrete in compression can be anticipated as follows:


𝜖𝑐 × 𝑛
𝜎𝑐 = 𝑓𝑐𝑚 𝜀𝑐 𝑛𝑘 (21)
𝜀𝑐𝑚 𝑛 ― 1 + (𝜀 )
𝑐𝑚

where

‒ fcm = peak stress;


‒ εcm = strain at peak stress;
‒ n = 0.8 + (fcm/17); and
‒ k = 0.67 + (fcm/62) when ɛc/ɛcm>1 or 1.0 when ɛc/ɛcm≤1.

5.6 Rate of strength development

The strength development rate and the chemical reaction of GeoPC are affected by a number of

factors depending on the chemical composition of source materials, the alkaline activators

(KOH, Na2SiO3 and NaOH), the curing conditions, and the mineralogical phases (Diaz et al.,

2010). A rise in temperature corresponds to a rise in the strength development rate, while the

ratio of alkaline liquid to binders does not influence the GeoPC (Chithra and Dhinakaran, 2014;

Gjorv and Sakai., 1999; D Hardjito et al., 2005; Hemalatha and Ramaswamy, 2017). In general,

the strength development of GeoPC becomes steady after 28 days (Kabir et al., 2015). In order to

attain an appropriate chemical composition in the growth of geopolymers, the favored procedure

is to mix FA with a great silica material. Reportedly, the mechanical properties of GeoPC are

55
Journal Pre-proof

intensely influenced by the connection between aggregate and cement paste at the interfacial

transition zone (ITZ) (Brough and Atkinson, 2000). Thus, it is found that the strength

development and ITZ in GeoPC improved through the creation of dense ITZ between the binder

matrix and aggregate at a greater SP amount (Mazloom et al., 2004). Several researchers have

observed that an FA-based GeoPC specimen shows higher compressive strength when cured at

high temperatures instead of ambient temperatures (Satpute Manesh et al., 2012; Thampi et al.,

2014). For instance, a 53% rise in the strength of GeoPC was reported after the

geopolymer was smoked using heat. Similarly, another study revealed an increment of

strength between 80% and 60% at 7-days and 28-days for ambient and oven-cured

samples, respectively. Also, Manesh et al. (Satpute Manesh et al., 2012) reported that the

strength development rate of concrete remains constant for up to 16 hours at 600 °C and 900 °C.

Higher than this temperature, concrete strength continues to increase at a lower rate, while at

1200 °C, the strength of concrete remains constant at all periods. GeoPC can rapidly gain

strength after 6 hours of heating at 1200 °C and shows an 80% strength gain after 24 hours. Such

strength development can be mainly attributed to the development of C–S–H cementitious gel on

the pore space and can improve the density of the resultant geopolymer binder matrix (Nath and

Kumar, 2013). It is also reported that the higher development of strength gained in external

contact curing was because of the increase in polarization of OH¯ to break Si-O and Al-O bonds

on the FA surface. Meanwhile, it is determined that the greater the Si content of the samples

cured at 60°C, the greater the strength achieved. Likewise, the strength of GeoPC reduces by

about 15-23% with a rise in the water to geopolymer solids ratio by mass (Calderón-Moreno et

al., 2002; D Hardjito et al., 2005; Zhuang et al., 2016b). Reportedly, the use of 100% GGBS-

based GeoPC mixes were shown to increase the strength by 41% between the results recorded at

56
Journal Pre-proof

7 days and 112 days at ambient room temperature curing and a 78% improvement in strength of

RHA-based GeoPC, depending on the RHA fineness, the ratio of FA/RHA, and sodium silicate

to NaOH (Ajay et al., 2012; B. N. Sangeetha, 2015; He et al., 2013; Mehta, 1977; Naji et al.,

2010; Shalini et al., 2016). Moreover, the RM-based GeoPC led to increases in strength

depending on the red mud and alkaline liquid contents; however, no substantial increase in

strength was observed with the addition of more than 30% RM (Dharmendra S. Ravat and Dave,

2017; Kumar and Kumar, 2013; PARAMGURU et al., 2004; Ye et al., 2016). The properties of

GeoPC–raw-materials (binders), additives, producing techniques, and chemical compositions

were seen to have a positive influence on the strength development of GeoPCs in both the short-

term and log-term.

6 Physical properties

GeoPC’s physical properties are influenced by many factors due to its mixed proportion of

binder, sand, aggregates, water, and alkaline liquids. As GeoPC matures, it continues to shrink

depending on the design density because of the ongoing reaction in the material. However, the

rate of shrinkage decreases rapidly and continues to reduce with time. The physical properties of

GeoPC, including its density shrinkage, porosity, and sorptivity, are described in the following

subsections.

6.1 Density

GeoPC has an average density (ASTM C 567) of 2020 kg/m3 to 2700 kg/m3 (A. Castel, 2016)

depending on the components in the structure. The amount of aggregate in GeoPC has the main

influence on its density, particularly the content of fine aggregate which improved the density

when greater volume was included in the matrix. The results of the slump test for GeoPC are

within the 80 mm ± 20 mm criterion, which is not only reliant on rheology but also on the

57
Journal Pre-proof

aggregate density (Nuaklong et al., 2016) and the density of the fibers (Kamseu et al., 2016; Lee

et al., 2016; Mazaheripour et al., 2011). The GGBS-based GeoPC with fine particles provides

density improvements of around 7.5% in comparison with POFA-based GeoPC; whereas GGBS-

based and FA-based GeoPCs densities are similar to that of normal concrete. It was observed that

the mortar density varies between 2014 kg/m3 and 2163 kg/m3. The combined aggregates mass is

taken in the range between 75% and 80% of the concrete mass (Castel and Foster, 2015; D

Hardjito et al., 2005; KThu and Murthy, 2015; Shaikh, 2016; Shaikh et al., 2015) because the

fine aggregates are limited to 30% of the total aggregates (KThu and Murthy, 2015;

Sreenivasulu, C. and Jawahar, 2015) and the addition of silicate solution increases the relative

density from 79% to 93% (He et al., 2010). When added to OPC concrete mixes, the Bayer-

derived geopolymer mortar aggregate shows a 30% lower density and a 50% higher compressive

strength (Bakharev, 2005). The reduction in the compressive strength of geopolymer mortars

with up to 10% of dry wastepaper sludge can be attributed to the existence of surfactants

(dissolved lignin residues) in the sludge; such decrease can cause the production of low-density

geopolymer mortar (A. Castel, 2016; Yan and Sagoe-Crentsil, 2012). The spherical particles (Eq.

11) of all materials can lead to a higher packing density compared with the crushed particles in a

wet state; therefore, the water retention and water demand in the spherical case are lesser than

those in the crushed case (Sakai, 1997). To increase the actual packing density, the grading span

must be increased by using a certain amount of finer particles, utilizing compact or rounded

particles, and continuous grading with increased ratio of fractions of larger particle size in

GeoPC (Görhan and Kürklü, 2014). Such effect is accredited to the fact that the delayed

pozzolanic reaction of FA increases the compressive strength and density of concrete over time

and ultimately leads to a carbonation resistance that is higher than that obtained in the

58
Journal Pre-proof

accelerated test (A. Castel, 2016). The density decreases as the proportion of POFA increases

(Kabir et al., 2015). Furthermore, RHA can reduce the bulk density (Ajay et al., 2012; Chennur

Jithendra, 2017; Provis and Deventer, 2009), and the small particle size of FA, which has a D50

of 7.6 μm, can increase the density of GeoPC concrete (Zhang et al., 2016). Some studies have

applied the particle technology in manufacturing OPC to enhance the granular distribution of

geopolymer materials and to attain high packing density, which in turn can help reduce the

required amount of active binders and alkaline activators (Ismail et al., 2011). Thus, compared to

the mortar with POFA, GGBS with fine particles can increase the density of GeoPC by

approximately 7.5% (Islam et al., 2014). Wardhono et al. (Wardhono et al., 2017) found that

increasing the packing density of the Al–Si gel matrix can positively affect the elastic modulus

and strength development of FAGP concrete between 90 days and 540 days. These results

altogether imply that reducing the concrete density can also reduce the transport cost, vehicle

wear, and road deformation (A. Castel, 2016). Furthermore, noteworthy cost savings are reported

because of density decrease as POFA concrete has almost 17–25% lower density than ordinary

concrete. It is also found that metakaolin-based GeoPC can achieve higher strength and

microstructure; therefore, it is applied for the formation of the geopolymeric gel of 1.45 g/cm3

density (Kong et al., 2007; Sakulich, 2011). Also, it was revealed that SFA-based GeoPC with

the addition of aluminium particles have led to a reduction in the bulk density by about 15.5%. In

general, the average density of GeoPC is equivalent to that of normal concrete, depending on its

composition and the underestimation integral to laser particle analysis where irregular particles

are presumed to be impeccably spherical (Eq. 22) (Shi et al., 2011).


0.6
𝐷
Geometric specific surface area (SSA) = 𝑃𝑎𝑟𝑡𝑖𝑐𝑙𝑒 𝑑𝑒𝑛𝑠𝑖𝑡𝑦 (22)

Where, D = size of particle, μm.

59
Journal Pre-proof

6.2 Dry shrinkage

Dry shrinkage (DS) (ASTM C 596) refers to the reduction of volume during the drying and

hardening processes (A. Castel, 2016). The shrinkage of GeoPC up to the age of 6 months was

determined to be similar to that of normal concrete of comparable strength (Duan et al., 2016).

The strains of FA-based GeoPCs depend significantly on the period of exposure in sulphate

solution of variable concentrations (Satpute Manesh et al., 2012). If the formation of these strains

is prevented, then the concrete faces tensile stress and develops cracks. Increasing the proportion

of slags above 10% can result in cracking due to DS (Sakulich, 2011). This phenomenon can be

reduced and controlled by adding fibers to the mix design of GeoPC (Rickard et al., 2014). A

concrete made of lightweight aggregates (e.g., RCA (Kovler and Roussel, 2011)) is known to

have higher DS (up to 50%) compared with conventional concrete (Mo et al., 2015b). Autoclave

curing can effectively reduce the DS of AAS (A. Castel, 2016). Yan and Crentsil monitored the

DS behavior of dry wastepaper sludge FA-added geopolymer mortars for up to 91 days (Yan and

Sagoe-Crentsil, 2012). Heat-cured FA-based GeoPC shows a very low DS of about 100 micro

strains after a year (D Hardjito et al., 2005). The addition of up to 10% dry wastepaper sludge

can reduce the DS of the resultant geopolymer mortars (Görhan and Kürklü, 2014; D Hardjito et

al., 2005; D. Hardjito et al., 2005; Hardjito et al., 2008; Noushini et al., 2016; Ryu et al., 2013).

Chindaprasirt et al. (Chindaprasirt et al., 2010) reported that using high-calcium FA as a

geopolymer source material can improve the DS of geopolymer mortars. They also found that

high-calcium geopolymer mortars demonstrate ten times improvement in terms of DS compared

with OPC mortars, thereby demonstrating the extraordinary dimensional stability of high-

calcium FA geopolymers (A. Castel, 2016). From the results of a moisture loss analysis, the

growth of cellulose fibers in the presence of moisture in GeoPC can increase the degree of DS,

60
Journal Pre-proof

while a continued curing at 70 °C for 7 to 28 days can reduce the strength of mixes with more

than 20% GGBS (Soutsos et al., 2016). The DS (0.025%) of GeoPC concrete becomes lower

than that of OPC concrete (0.09%) after 12 weeks (Singh et al., 2015) and the incorporation of

nano-TiO2 particles refines the microstructure and lowers the DS of GeoPC (Duan et al., 2016).

Another study of 20% FA-based– and GGBS-based GeoPC with sodium silicate to sodium

hydroxide (SS/SH) ratio revealed that shrinkage reduced with the rise of slag content and

reduction in SS/SH ratio in GeoPC cured at room temperature, leading it to be comparable to that

of normal concrete of similar strength (Deb et al., 2015). Meanwhile, the use of 50% RM in

designing GeoPC is seen to greatly increase the strength; however, exceeding the 50%

percentage can cause a lot of shrinkage cracks (Dharmendra S. Ravat and Dave, 2017; He et al.,

2012; Kumar and Kumar, 2013; PARAMGURU et al., 2004). Furthermore, the low drying

shrinkage of GeoPC provides support to the long-term performance of GeoPC elements, leading

to use in infrastructure applications.

6.3 Porosity

Porosity (ASTM C 830) which is usually caused by several related factors, such as relative

humidity, degree of reaction, and macroscopic properties of GeoPC, is vital to the development

of GeoPC (Noushini et al., 2016; Provis et al., 2005). Pore size distribution and porosity are the

most important characteristics to be analysed in order to study corrosion in GeoPC. Porosity of

GeoPC is highly influenced by the volume of binder added. For example, in the design of RHA-

based GeoPC, an increase in the amount of RHA necessitates an upsurge in water/cement ratio.

Due to that, RHA is a greatly porous material (B. N. Sangeetha, 2015; Habeeb and Mahmud,

2010; Naji et al., 2010). Baltazar et al. (Baltazar et al., 2014) found that the effectiveness of

surface treatment is more significant in concrete with higher porosity because treatment agents

61
Journal Pre-proof

can penetrate more easily and deeply into this concrete. Those geopolymeric mortars with

sodium hydroxide molarities of 14 M and 18 M can absorb around 20% of water because of their

open porosity (A. Castel, 2016). In alkali-activated cement (AAC), the curve of 100% FA binder

has presented dominant pore diameters of 10–100 nm in total porosity of 20% geopolymer

binders (Diaz et al., 2010; Gunasekara et al., 2016; Hemalatha and Ramaswamy, 2017). In this

binder, the maximal diameters distribution was around 27 nm and the pores in the mesopores

interval, 10–50 nm, accounted for 80% frequency of total porosity. This curve differs from the

bimodal profile of the pore size distributions in alkali-activated FA that is cured for 28 days,

where the pores are located at 100 nm and 1000 nm (Khale and Chaudhary, 2007; Ma et al.,

2013). Increasing the number of cotton fabric layers also gradually increases the porosity of

composites by 20% to 30% (Yan et al., 2016). Provis et al. (Provis et al., 2012) observed that

adding GGBS not only reduces porosity but also produces a pore refinement effect. This finding

indicates that the excess GGBS can increase the porosity of concrete (KThu and Murthy, 2015;

Mo et al., 2016; Sreenivasulu, C. and Jawahar, 2015). Porosity and cracks also have harmful

effects on the modulus of elasticity of RCA (Zhang and Bentley, 2003). Moreover, the use of

fine RCA is categorized by a greater porosity than coarse ones because of the volume of

enclosed cement mortar. Reportedly, the use of raw RHA particle size in GeoPC at 1000º C

heating displayed 87% porosity and 450 m2/ g SSA but the larger particles improved strength

(Ajay et al., 2012; Chu et al., 2016; Provis and Deventer, 2009). As mentioned earlier, adding

slag increases the workability of the matrix because this material needs less liquid compared with

metakaolin for particle wetting, which in turn can help reduce porosity (Sakulich, 2011).

Reducing the alkalinity of the binder matrix and adding aggregates and Si carbide sludge can

also help reduce porosity (Almeida et al., 2013; Druta, 2003; Nuaklong et al., 2016; Pouhet and

62
Journal Pre-proof

Cyr, 2016; Prud’homme et al., 2015). Reportedly, the addition of SF by up to 5% improved the

strength of GeoPC, but additional increase in SF initiated a reduction in strength, contributed to

better microstructure and revealed lesser porosity (Atiş et al., 2005; Bhavsar et al., 2014;

Mazloom et al., 2004; Memon et al., 2013). Meanwhile, the 40% POFA-based GeoPC with 2.5%

NaOH showed an increase in strength by 95 % and a reduction in porosity by 5.27% at 28 days

(Kabir et al., 2015; Salih et al., 2014). Another research revealed that the addition of less than

30% RM in GeoPC caused higher porosity performance. In GeoPC, porosity is generally

affected by the pores formed with the diameter size depending on the volume of binder included,

in particular, when the binder added with the volume was more than half of OPC.

6.4 Sorptivity

Sorptivity (ASTM C1585) refers to the capability of concrete to absorb and transmit water by

means of capillary suctions (Davidovits, 1999; Mo et al., 2016; Rickard et al., 2014). The

sorptivity of geopolymers (5 - 30 𝜇m/s1/2) greatly depend on their water content and forming

pressure (A. Castel, 2016; Nuaklong et al., 2016). The sorptivity of GeoPC can increase with a

rise in grade; for instance, the increase in FA content can lead to improvements in sorptivity

(Thokchom et al., 2009). The defiance to the sorptivity of GeoPCs is inversely balanced to the

water/binder ratio, fineness of particles, and extra water (Deb et al., 2016; Thokchom et al.,

2009). The initial sorptivity of GeoPC is measured during the first six hours of water absorption

(Shaikh, 2016). The water sorptivity of 100 mm GeoPC specimens is usually measured at 28

days (Soutsos et al., 2016). The measurement begins by placing the specimens inside an oven

and drying them at 105 °C ± 5 °C for 48 h until a constant weight is recorded. These specimens

are then greased at all sides and covered with cling film in order for the water to be absorbed

vertically when these specimens are placed 5 mm under the water surface (Fig. 15) (Hadjsadok et

63
Journal Pre-proof

al., 2012). Reportedly, the 25% replacement of natural aggregate by RCA in GeoPC presented

better sorptivity than normal RAC. Another research reported that the POFA-based GeoPC

revealed a lower sorptivity due to contribution to the rise of particles’ surface area (Detphan and

Chindaprasirt, 2009; Somna et al., 2011; Wongpa et al., 2010).

Fig. 15: Setup for the sorptivity test of GeoPC (Hadjsadok et al., 2012)

Meanwhile, the use of 50% GGBS and 50% FA in GeoPC can reduce the sorptivity in GeoPC

with about 36% compared to normal concrete (Thokchom et al., 2009) Afterwards, the

specimens were weighed at 1, 4, 9, 16, 25, 36, 49, and 64 min to determine their weight gain.

Those GeoPC specimens with lower alkali content have higher water sorptivity (Thokchom et

al., 2009). Meanwhile, increasing the metakaolin content and activator concentration as well as

adding RCA, GGBS, 2% nano-silica dosage, metakaolin, FA, and 8% Na2O can reduce the water

sorptivity of GeoPC by 20% (Bernal et al., 2012; Davidovits, 1999; Deb et al., 2016; Mo et al.,

2016; Shaikh, 2016; Singh and Siddique, 2015; Thokchom et al., 2009; Wardhono et al., 2017).

Increasing the flowability property of GeoPC can also lower the porosity of concrete and

subsequently lessen its absorption of water (Chi and Huang, 2014), However, the sorptivity of

64
Journal Pre-proof

GeoPC exhibited lower rate when compared to normal concrete of different grades. The

sorptivity of the concrete specimens is calculated using Eqn. (23). The same experimental setup

has been adopted in other studies (Hadjsadok et al., 2012; Liu et al., 2016)

i = A + St1/2 (23)

where; – S = sorptivity coefficient, mm/min1/2;


‒ i = cumulative volume of water absorbed at time, mm/mm2;
‒ A = surface area of the test specimen, mm2; and
‒ t = recorded time, min.
7 Conclusion

GeoPC is an inventive construction material that provides an alternative to OPC, and it is made

by the chemical activity of inorganic particles of waste materials. GeoPC has become more

common in past decades due to being more environmentally friendly as opposed to conventional

OPC. Further, economic and environmental reasons necessitate the amendment of current

concrete production materials (OPC). This can be achieved through the clean production and use

of GeoPC. It is found that the production of OPC and GeoPC using secondary industrial raw

waste materials, such as FA, SFA, GGBS, and RM, is a better substitute to traditional OPC, and

because of that, GeoPC can provide ultra-high early strength, greater durability, improved

economic benefits, less CO2 emissions during production, lower use of sodium silicate solution,

and lengthier service life in a number of RC applications, in particular, for use in transportation

infrastructure construction. Thus, the production of industrial-by-products-based geopolymer

greatly depends on alkali activated geopolymerization occurring under moderate conditions and

is deemed as a cleaner procedure because of the higher reduction of CO2 emissions during

manufacturing when compared to OPC. It is shown that the production of GeoPC needs intensive

care and exact material composition. During the activation process while producing the

geopolymer, great alkalinity also obliges a safety danger and improved energy consumption and

65
Journal Pre-proof

production of greenhouse gases. Moreover, GeoPC is highly influenced by the curing

temperature and time as well as by the properties and proportions of the constituent materials. In

this study, it was revealed that GeoPC has excellent compressive strength and this led GeoPC to

be identified as a high potential product used in the fabrication of several structural concrete

applications. It was also determined that GeoPC has a significant resistance to acid, excellent

resistance to sulfate attack, experiences low creep, and slightly suffers from drying shrinkage.

The outstanding factors that affect the properties of the fresh and hardened GeoPC have been

discussed. Based on this review study, it was observed that the vast majority of previous research

work focused on the specific properties of geopolymer, such as compressive strength, rather than

focusing on their characteristics such as alkaline activator solutions. Also, components such as

natural and artificial fibers in GeoPC and their effect on the strength of the GeoPC have not

received significant attention over the years. Therefore, based on this review, the following

conclusions are drawn:

 Binder paste containing small diameter carbon nano-fiber demonstrates high sensitivity

and stable sensing properties in GeoPC.

 Gel nuclei particles must be sufficiently stable to resist depolymerization and in order to

begin a new gel phase that will be primarily responsible for strength and durability

enhancement of GeoPC.

 Silicate and Aluminate monomers that condense stabilization mechanism in a GeoPC

network are considered among the most important components of concrete that guarantee

fast chemical reaction of Si–Al minerals under alkaline conditions.

 The desired density and strength of concrete depends on the method of proportioning

materials, design codes and construction guidelines.

66
Journal Pre-proof

 The stability of GeoPC production depends on numerous factors such as curing

temperature, setting and curing time, molarity of alkaline activator and mix ratio.

 The properties of GeoPC in the fresh state determine various aspects of workability that

control segregation resistance, passing ability and filling ability of GeoPC which must be

cautiously controlled.

 GeoPC exhibits high early strength and has been effectively utilized in precast industries.

This is demonstrated by its ability to produce large-scale GeoPC structure within a short

duration with minimal tendency to breakage during transportation.

With this in mind, it can be said that GeoPC is a superior alternative material to cement and can

be effectively used to replace OPC for practical use in construction industries worldwide. Future

studies should focus on the improvement of strength and durability of GeoPC in hardened state

through inclusion of fibers. Furthermore, future studies should also consider investigating the

influence of aggregate content and additives on the engineering properties of GeoPC. Also,

suitable guidelines for selection of aggregate contents in GeoPC should be developed with clear

mix design procedure. Investigations on the porosity property of geopolymer foams prepared by

using mixed-foaming method should be carried out. The properties of these foams are no better

than those of traditional porous materials, such as glass foam or autoclaved-aerated concrete.

These issues limit the potential use of geopolymer foams as thermal insulation materials and

prevent them from competing with traditional porous inorganic materials. Lastly, further studies

are required to compare the cost of GeoPC to that of conventional concrete.

8 Acknowledgment

The authors gratefully acknowledge the financial support by the Department of Civil

Engineering, College of Engineering, Prince Sattam Bin Abdulaziz University, Saudi Arabia;

67
Journal Pre-proof

and the Department of Civil Engineering, Faculty of Engineering and IT, Amran University,

Yemen, for this research.

9 References

A. Castel, 2016. Bond between steel reinforced and Geopolymer concrete Ch14.pdf. Handb. low
carbon Concr.
ACI 233R-95 Committee Report, 1997. GGBFS as a Cementitious Constituents in Concrete,
ACI Manual of Concrete Practice, Part I.
Adam, A.., 2009. Strength and durability properties of alkali activated slag and fly ash-based
geopolymer concrete.
Adams, T.H., 2017. American Coal Ash Association Production and Use News Release. Am.
Coal Ash Assoc.
Adewuyi, A.P., Sulaiman, I.A., Akinyele, J.O., 2017. Compressive Strength and Abrasion
Resistance of Concretes under Varying Exposure Conditions. Open J. Civ. Eng.
https://doi.org/10.4236/ojce.2017.71005
Aggarwal, P., Siddique, R., Aggarwal, Y., Gupta, S.M., 2008. Self-compacting concrete-
procedure for mix design. Leonardo Electron. J. Pract. Technol. 12, 15–24.
Aharon-Shalom, E., Heller, a., 1982. Tensile Strength and Bonding Characteristics of Self-
Compacting Concrete. J. Electrochem. Soc.
Ahmari, S., Zhang, L., 2012. Production of eco-friendly bricks from copper mine tailings
through geopolymerization. Constr. Build. Mater. 29, 323–331.
https://doi.org/10.1016/j.conbuildmat.2011.10.048
Ajay, K., Mohanta, K., Kumar, D., Parkash, O., 2012. Properties and industrial applications of
rice husk: a review. Int. J. Emerg. Technol. Adv. Eng. 2, 86–90.
Al-Majidi, M.H., Lampropoulos, A., Cundy, A.B., 2017. Tensile properties of a novel fibre
reinforced geopolymer composite with enhanced strain hardening characteristics. Compos.
Struct. 168, 402–427. https://doi.org/10.1016/j.compstruct.2017.01.085
Al-Mulla, I.F., 2010. Enhance concrete performance using natural materials as partial
replacement of cement. Acad. Sci. J. 18.
Al-Qadri, F.A., Saad, A., Aldlaee, A.A., 2009. Effect of some admixtures on heat of hydration
reaction of cement pastes produced in Yemen, Saudi Arabia, and Egypt. J. Eng. Sci. 27.
Alam, S.Y., Saliba, J., Loukili, A., 2014. Fracture examination in concrete through combined
digital image correlation and acoustic emission techniques. Constr. Build. Mater. 69, 232–
242. https://doi.org/10.1016/j.conbuildmat.2014.07.044
Aldea, C.-M., Young, F., Wang, K., Shah, S.P., 2000. Effects of curing conditions on properties
of concrete using slag replacement. Cem. Concr. Res. 30, 465–472.
https://doi.org/10.1016/S0008-8846(00)00200-3
Ali, M., Zurisman, A., 2015. Performance of geopolymer concrete in fire. Swinburne University
of Technology.
Almeida, A.E.F.S., Tonoli, G.H.D., Santos, S.F., Savastano, H., 2013. Improved durability of
vegetable fiber reinforced cement composite subject to accelerated carbonation at early age.
Cem. Concr. Compos. 42, 49–58. https://doi.org/10.1016/j.cemconcomp.2013.05.001
Alomayri, T., Shaikh, F.U.A., Low, I.M., 2014a. Effect of fabric orientation on mechanical
properties of cotton fabric reinforced geopolymer composites. Mater. Des. 57, 360–365.

68
Journal Pre-proof

https://doi.org/10.1016/j.matdes.2014.01.036
Alomayri, T., Shaikh, F.U.A., Low, I.M., 2014b. Synthesis and mechanical properties of cotton
fabric reinforced geopolymer composites. Compos. Part B Eng. 60, 36–42.
https://doi.org/10.1016/j.compositesb.2013.12.036
American Concrete Institute., A., Malhotra, V.M., 2000. High-volume fly ash system: the
concrete solution for sustainable development, ACI Materials Journal. American Concrete
Institute.
Ananthayya, M., WP., P.K., 2014. Effect of Partial Replacement of Sand by Iron Ore Tailing
(IOT) and Cement by Ground Granulated Blast Furnace Slag (GGBFS) on the Compressive
Strength of Concrete. Int. J. Eng. 3.
Anuar, K., Ridzuan, A., Ismail, S., 2011. Strength characteristics of geopolymer concrete
containing recycled concrete aggregate. Int. J. Civ. Environ. Eng. 11, 59–62.
Anuradha, R., Thirumala, R., John, P., 2014. Optimization of molarity on workable self-
compacting geopolymer concrete and strength study on SCGC by replacing fly ash with
silica fume and GGBFS. Int J Adv Struct Geotech Eng.
AS, A.S., 2009. Concrete structures. AS3600-2001. Sydney (Australia).
ASTM, 2012. Standard Specification for Silica Fume Used in Cementitious Mixtures. Astm
C1240 1–7. https://doi.org/10.1520/C1240-14.2
ASTMC, 2004. Standard Test Method for Density, Relative Density (Specific Gravity), and
Absorption of Coarse Aggregate, Annual book of ASTM standards. West Conshohocken.
ASTMC115-96, 1996. Standard Test Method for Fineness of Portland Cement by the
Turbidimeter.
ASTMC666, 1997. Standard Test Method for Resistance of Concrete to Rapid Freezing and
Thawing.
ASTME119, 2012. Standard test methods for fire tests of building construction and materials.
West Conshohocken, PA.
Atiş, C.D., Özcan, F., Kılıç, A., Karahan, O., Bilim, C., Severcan, M.H., 2005. Influence of dry
and wet curing conditions on compressive strength of silica fume concrete. Build. Environ.
40, 1678–1683. https://doi.org/10.1016/j.buildenv.2004.12.005
Azreen, N.M., Rashid, R.S.M., Haniza, M., Voo, Y.L., Mugahed Amran, Y.H., 2018. Radiation
shielding of ultra-high-performance concrete with silica sand, amang and lead glass. Constr.
Build. Mater. 172, 370–377. https://doi.org/10.1016/j.conbuildmat.2018.03.243
B. N. Sangeetha, 2015. Effect of Rice Husk Ash and GGBS on Performance of Concrete. Int. J.
Eng. Res. V4, 491–495. https://doi.org/10.17577/ijertv4is120515
Bakharev, T., 2005. Resistance of geopolymer materials to acid attack. Cem. Concr. Res.
https://doi.org/10.1016/j.cemconres.2004.06.005
Bakri, A., Mustafa, M., Mohammed, H., Kamarudin, H., Niza, I.K., Zarina, Y., 2011. Review on
fly ash-based geopolymer concrete without Portland Cement. J. Eng. Technol. Res. 3, 1–4.
Bakri, A.M.M.A., Kamarudin, H., Bnhussain, M., Nizar, I.K., Mastura, W.I.W., 2011.
Mechanism and Chemical Reaction of Fly Ash Geopolymer Cement- A Review.
Bakri, M. Al, A. M., H.K., Bnhussain, M., Rafiza, A.R., Zarina., Y., 2012. Effect of Na 2 SiO
3/NaOH Ratios and NaOH Molarities on Compressive Strength of Fly-Ash-Based
Geopolymer. ACI Mater. J. 109.
Baltazar, L., Santana, J., Lopes, B., Paula Rodrigues, M., Correia, J.R., 2014. Surface skin
protection of concrete with silicate-based impregnations: Influence of the substrate
roughness and moisture. Constr. Build. Mater. 70, 191–200.

69
Journal Pre-proof

https://doi.org/10.1016/j.conbuildmat.2014.07.071
Basha S, M., Reddy Ch, B., K, V., 2016. Strength behaviour of geopolymer concrete replacing
fine aggregates by M- sand and E-waste. Int. J. Eng. Trends Technol. 40, 401–407.
https://doi.org/10.14445/22315381/ijett-v40p265
Bashar, I.I., Alengaram, U.J., Jumaat, M.Z., Islam, A., Santhi, H., Sharmin, A., 2016.
Engineering properties and fracture behaviour of high volume palm oil fuel ash based fibre
reinforced geopolymer concrete. Constr. Build. Mater.
https://doi.org/10.1016/j.conbuildmat.2016.02.022
Bernal, S.A., Mejía de Gutiérrez, R., Provis, J.L., 2012. Engineering and durability properties of
concretes based on alkali-activated granulated blast furnace slag/metakaolin blends. Constr.
Build. Mater. 33, 99–108. https://doi.org/10.1016/j.conbuildmat.2012.01.017
Bharath, K.N., Basavarajappa, S., 2014. Flammability Characteristics of Chemical Treated
Woven Natural Fabric Reinforced Phenol Formaldehyde Composites. Procedia Mater. Sci.
5, 1880–1886. https://doi.org/10.1016/j.mspro.2014.07.507
Bhavsar, G.D., Talavia, K.R., Amin, D.P.S.M.B., Parmar, A.A., 2014. Workability properties of
geopolymer concrete using accelerator and silica fume as an admixture. Int. J. Technol. Res.
Eng. 8.
Bhikshma, V., 2012. An experimental investigation on properties of geopolymer concrete (no
cement concrete). Asian J. Civ. Eng. Build. Hous. 13, 841–853.
Biswas, W.K., Cooling, D., 2013. Sustainability Assessment of Red Sand as a Substitute for
Virgin Sand and Crushed Limestone. J. Ind. Ecol. n/a-n/a.
https://doi.org/10.1111/jiec.12030
Board, N., 2012. Handbook on Agro Based Industries. Niir Proj. Consult. Serv.
Boonserm, K., Sata, V., Pimraksa, K., Chindaprasirt, P., 2012. Improved geopolymerization of
bottom ash by incorporating fly ash and using waste gypsum as additive. Cem. Concr.
Compos. 34, 819–824. https://doi.org/10.1016/j.cemconcomp.2012.04.001
Boyd, A.J., Mindess, S., Skalny, J., 2002. Cement and concrete-- trends and challenges.
American Ceramic Society, Westerville, OH.:
Brake, N.A., Allahdadi, H., Adam, F., 2016. Flexural strength and fracture size effects of
pervious concrete. Constr. Build. Mater. 113, 536–543.
https://doi.org/10.1016/j.conbuildmat.2016.03.045
Breña, S.F., Bramblett, R.M., Benouaich, M.A., Wood, S.L., Kreger., M.E., 2001. Use of carbon
fiber reinforced polymer composites to increase the flexural capacity of reinforced concrete
beams.
Brough, A.., Atkinson, A., 2000. Automated identification of the aggregate–paste interfacial
transition zone in mortars of silica sand with Portland or alkali-activated slag cement paste.
Cem. Concr. Res. 30, 849–854. https://doi.org/10.1016/S0008-8846(00)00254-4
Calderón-Moreno, J.M., Schehl, M., Popa, M., 2002. Superplastic behavior of zirconia-
reinforced alumina nanocomposites from powder alcoxide mixtures. Acta Mater. 50, 3973–
3983. https://doi.org/10.1016/S1359-6454(02)00163-5
Castel, 2016. Bond between steel reinforced and Geopolymer concrete Ch14.pdf. Handb. low
carbon Concr.
Castel, A., Foster, S.J., 2015. Bond strength between blended slag and Class F fly ash
geopolymer concrete with steel reinforcement. Cem. Concr. Res. 72, 48–53.
https://doi.org/10.1016/j.cemconres.2015.02.016
Chandra, S., Berntsson., L., 2002. Lightweight aggregate concrete. William Andrew Publishing,

70
Journal Pre-proof

Norwich, New York.


Chang, H.-L., Shih, W.-H., 2000. Synthesis of Zeolites A and X from Fly Ashes and Their Ion-
Exchange Behavior with Cobalt Ions. https://doi.org/10.1021/IE990860S
Chen, R., Ahmari, S., Zhang, L., 2014. Utilization of sweet sorghum fiber to reinforce fly ash-
based geopolymer. J. Mater. Sci. 49, 2548–2558. https://doi.org/10.1007/s10853-013-7950-
0
Chennur Jithendra, R., 2017. Geopolymer concrete with self-compacting: a review. Int. J. Civ.
Eng. Technol. 8.
Chi, M., Huang, R., 2014. Effect of circulating fluidized bed combustion ash on the properties of
roller compacted concrete. Cem. Concr. Compos. 45, 148–156.
https://doi.org/10.1016/j.cemconcomp.2013.10.001
Chindaprasirt, P., Chareerat, T., Hatanaka, S., Cao, T., 2010. High-Strength Geopolymer Using
Fine High-Calcium Fly Ash. J. Mater. Civ. Eng. https://doi.org/10.1061/(asce)mt.1943-
5533.0000161
Chindaprasirt, P., De Silva, P., Hanjitsuwan, S., 2014. Effect of High-Speed Mixing on
Properties of High Calcium Fly Ash Geopolymer Paste. Arab. J. Sci. Eng. 39, 6001–6007.
https://doi.org/10.1007/s13369-014-1217-1
Chindaprasirt, P., Rattanasak, U., Taebuanhuad, S., 2013. Resistance to acid and sulfate solutions
of microwave-assisted high calcium fly ash geopolymer. Mater. Struct. Constr.
https://doi.org/10.1617/s11527-012-9907-1
Chithra, S., Dhinakaran, G., 2014. Effect of hot water curing and hot air oven curing on admixed
concrete. Int. J. ChemTech Res. 6, 1516–1523.
Chu, H., Jiang, J., Sun, W., Zhang, M., 2016. Mechanical and physicochemical properties of
ferro-siliceous concrete subjected to elevated temperatures. Constr. Build. Mater. 122, 743–
752.
Collins, M.P., Mitchell, D., MacGregor, J.G., 1993. Structural design considerations for high-
strength concrete. Concr. Int. 15, 27–34.
Corinaldesi, V., Moriconi, G., 2009. Influence of mineral additions on the performance of 100%
recycled aggregate concrete. Constr. Build. Mater.
https://doi.org/10.1016/j.conbuildmat.2009.02.004
Davidovits, J., 2015. Geopolymer chemistry & applications.
Davidovits, J., 1999. Chemistry of geopolymeric systems, terminology, in: Second International
Conference on Geopolymer. pp. 9–40.
Davidovits, P.J., 2002. 30 Years of Successes and Failures in Geopolymer Applications . Market
Trends and Potential Breakthroughs . Geopolymer 2002 Conf.
https://doi.org/10.1017/CBO9781107415324.004
Deb, P.S., Nath, P., Sarker, P.K., 2015. Drying Shrinkage of Slag Blended Fly Ash Geopolymer
Concrete Cured at Room Temperature. Procedia Eng. 125, 594–600.
https://doi.org/10.1016/j.proeng.2015.11.066
Deb, P.S., Sarker, P.K., Barbhuiya, S., 2016. Sorptivity and acid resistance of ambient-cured
geopolymer mortars containing nano-silica. Cem. Concr. Compos.
https://doi.org/10.1016/j.cemconcomp.2016.06.017
Detphan, S., Chindaprasirt, P., 2009. Preparation of fly ash and rice husk ash geopolymer. Int. J.
Miner. Metall. Mater. 16, 720–726. https://doi.org/10.1016/S1674-4799(10)60019-2
Detwiler, R.., Bhatty, J.I., S. Battacharja, 1996. Supplementary cementing materials for use in
blended cements.

71
Journal Pre-proof

Dhakal, S., 2009. Urban energy use and carbon emissions from cities in China and policy
implications. Energy Policy. https://doi.org/10.1016/j.enpol.2009.05.020
Dharmendra S. Ravat, S.G.S., Dave, S. V, 2017. Utilization of Red Mud in Geopolymer
Concrete-A Review. Int. J. Adv. Eng. Res. 4.
Diaz, E.I., Allouche, E.N., Eklund, S., 2010. Factors affecting the suitability of fly ash as source
material for geopolymers. Fuel 89, 992–996. https://doi.org/10.1016/j.fuel.2009.09.012
Dimas, D., Giannopoulou, I., Panias, D., 2009. Polymerization in sodium silicate solutions: A
fundamental process in geopolymerization technology. J. Mater. Sci. 44, 3719–3730.
https://doi.org/10.1007/s10853-009-3497-5
Dodoo-Arhin, D., Nuamah, R.A., Agyei-Tuffour, B., Obada, D.O., Yaya, A., 2017. Awaso
bauxite red mud-cement based composites: Characterisation for pavement applications.
Case Stud. Constr. Mater. https://doi.org/10.1016/j.cscm.2017.05.003
Druta, C., 2003. Tensile strength and bonding characteristics of self-compacting concrete.
Polytechnic University of Bucharest.
Duan, P., Yan, C., Luo, W., Zhou, W., 2016. Effects of adding nano-TiO2 on compressive
strength, drying shrinkage, carbonation and microstructure of fluidized bed fly ash based
geopolymer paste. Constr. Build. Mater. https://doi.org/10.1016/j.conbuildmat.2015.12.095
EFNARC, F., 2002. Specification and guidelines for self-compacting concrete. European 29
Federation of National Associations Representing producers and applicators of specialist 30
building products for Concrete (EFNARC) 32, 31.
Esping, O., Löfgren., I., 2005. Investigation of early age deformation in self-compacting
concrete, in: The Knud Højgaard Conference on Advanced Cement-Based Materials-
Research and Teaching.
Etxeberria, M., Vázquez, E., Marí, A., Barra, M., 2007. Influence of amount of recycled coarse
aggregates and production process on properties of recycled aggregate concrete. Cem.
Concr. Res. 37, 735–742. https://doi.org/10.1016/j.cemconres.2007.02.002
European Commision, 2014. A policy framework for climate and energy in the period [WWW
Document]. Commun. FROM Comm. TO Eur. Parliam. Counc. Eur. Econ. Soc. Comm.
Comm. Reg.
Fernandez-Jimenez, A.M., Palomo, A., Lopez-Hombrados, C., 2006. Engineering Properties of
Alkali-activated fly ash concrete. ACI Mater. J. 103.
Ganesan, N., Abraham, R., Deepa Raj, S., Sasi, D., 2014. Stress–strain behaviour of confined
Geopolymer concrete. Constr. Build. Mater. 73, 326–331.
https://doi.org/10.1016/j.conbuildmat.2014.09.092
Ganesh Babu, K., Sree Rama Kumar, V., 2000. Efficiency of GGBS in concrete. Cem. Concr.
Res. 30, 1031–1036. https://doi.org/10.1016/S0008-8846(00)00271-4
Ganesh, M., M, V.S.A., Sangeetha., A., 2016. Effect of Different Types of Super Plasticizer on
Fresh and Hardened Properties of Self Consolidating Geopolymer Concrete. Int. J. earth
Sci. Eng. 9.
Ganeshan, ahima, Venkataraman, S., 2017. Self Consolidating Geopolymer Concrete as an Aid
to Green Technologies - Review on Present Status. Asian J. Res. Soc. Sci. Humanit.
https://doi.org/10.5958/2249-7315.2017.00187.3
Gaochuang, C., Noguchi, T., Degée, H., Zhao, J., Kitagaki, R., 2016. Volcano-related materials
in concretes: a comprehensive review. Environ. Sci. Pollut. Res. 23, 7220–7243.
Geopolymer, J.D. of, 1988, undefined, n.d. Soft mineralurgy and geopolymers. researchgate.net.
Gjorv, O.E., Sakai., K., 1999. Concrete technology for a sustainable development in the 21st

72
Journal Pre-proof

century. CRC Press.


Gonçalves, M.R.F., Bergmann, C.P., 2007. Thermal insulators made with rice husk ashes:
Production and correlation between properties and microstructure. Constr. Build. Mater. 21,
2059–2065. https://doi.org/10.1016/j.conbuildmat.2006.05.057
Görhan, G., Kürklü, G., 2014. The influence of the NaOH solution on the properties of the fly
ash-based geopolymer mortar cured at different temperatures. Compos. Part B Eng. 58,
371–377. https://doi.org/10.1016/j.compositesb.2013.10.082
Gunasekara, C., Law, D.W., Setunge, S., 2016. Long term permeation properties of different fly
ash geopolymer concretes. Constr. Build. Mater. 124, 352–362.
https://doi.org/10.1016/j.conbuildmat.2016.07.121
Gursel, A.P., Maryman, H., Ostertag, C., 2016. A life-cycle approach to environmental,
mechanical, and durability properties of “green” concrete mixes with rice husk ash. J.
Clean. Prod. https://doi.org/10.1016/j.jclepro.2015.06.029
Habeeb, G.A., Mahmud, H. Bin, 2010. Study on properties of rice husk ash and its use as cement
replacement material. Mater. Res. 13, 185–190. https://doi.org/10.1590/s1516-
14392010000200011
Hadjsadok, A., Kenai, S., Courard, L., Michel, F., Khatib, J., 2012. Durability of mortar and
concretes containing slag with low hydraulic activity. Cem. Concr. Compos. 34, 671–677.
https://doi.org/10.1016/j.cemconcomp.2012.02.011
Haider, G.M., Sanjayan, J.G., Ranjith, P.G., 2014. Complete triaxial stress–strain curves for
geopolymer. Constr. Build. Mater. 69, 196–202.
https://doi.org/10.1016/j.conbuildmat.2014.07.058
Han, B., Yu, X., Ou, J., 2014. Self-Sensing Concrete in Smart Structures, Self-Sensing Concrete
in Smart Structures. https://doi.org/10.1016/C2013-0-14456-X
Hardjito, D., Cheak, C.C., Lee Ing, C.H., 2008. Strength and Setting Times of Low Calcium Fly
Ash-based Geopolymer Mortar. Mod. Appl. Sci. 2. https://doi.org/10.5539/mas.v2n4p3
Hardjito, D., Wallah, S.E., Sumajouw, D.M.J., Rangan., B. V., 2005. Introducing fly ash-based
geopolymer concrete: manufacture and engineering properties, in: The 30th Conference on
Our World in Concrete and Structures.
Hardjito, D, Wallah, S.E., Sumajouw, D.M.J., Rangan, B. V, 2005. Fly Ash-Based Geopolymer
Concrete. Aust. J. Struct. Eng. 6, 77–86. https://doi.org/10.1080/13287982.2005.11464946
Hassan, A., Arif, M., Shariq, M., 2019. Use of geopolymer concrete for a cleaner and sustainable
environment – A review of mechanical properties and microstructure. J. Clean. Prod.
https://doi.org/10.1016/j.jclepro.2019.03.051
Hawes, D.W., Banu, D., Feldman, D., 1992. The stability of phase change materials in concrete.
Sol. Energy Mater. Sol. Cells 27, 103–118. https://doi.org/10.1016/0927-0248(92)90113-4
Hawes, D.W., Banu, D., Feldman, D., 1990. Latent heat storage in concrete. II. Sol. Energy
Mater. 21, 61–80. https://doi.org/10.1016/0165-1633(90)90043-Z
He, J., Jie, Y., Zhang, J., Yu, Y., Zhang, G., 2013. Synthesis and characterization of red mud and
rice husk ash-based geopolymer composites. Cem. Concr. Compos. 37, 108–118.
https://doi.org/10.1016/j.cemconcomp.2012.11.010
He, J., Zhang, J., Yu, Y., Zhang, G., 2012. The strength and microstructure of two geopolymers
derived from metakaolin and red mud-fly ash admixture: A comparative study. Constr.
Build. Mater. 30, 80–91. https://doi.org/10.1016/j.conbuildmat.2011.12.011
He, P., Jia, D., Lin, T., Wang, M., Zhou, Y., 2010. Effects of high-temperature heat treatment on
the mechanical properties of unidirectional carbon fiber reinforced geopolymer composites.

73
Journal Pre-proof

Ceram. Int. 36, 1447–1453. https://doi.org/10.1016/j.ceramint.2010.02.012


Hein, M., Arena, S., 2010. Foundations of college chemistry, alternate. John Wiley & Sons.
Hemalatha, T., Ramaswamy, A., 2017. A review on fly ash characteristics – Towards promoting
high volume utilization in developing sustainable concrete. J. Clean. Prod. 147, 546–559.
https://doi.org/10.1016/j.jclepro.2017.01.114
Huseien, G.F., Al-Fasih, M.Y., Hamzah, H.K., 2015. Performance of Self-Compacting Concrete
With Different Sizes of Recycled Ceramic Aggregates. Int. J. Innov. Res. Creat. Technol. 1,
264–269.
Islam, A., Alengaram, U.J., Jumaat, M.Z., Bashar, I.I., 2014. The development of compressive
strength of ground granulated blast furnace slag-palm oil fuel ash-fly ash based geopolymer
mortar. Mater. Des. 56, 833–841. https://doi.org/10.1016/j.matdes.2013.11.080
Islam, A., Alengaram, U.J., Jumaat, M.Z., Bashar, I.I., Kabir, S.M.A., 2015. Engineering
properties and carbon footprint of ground granulated blast-furnace slag-palm oil fuel ash-
based structural geopolymer concrete. Constr. Build. Mater. 101, 503–521.
https://doi.org/10.1016/j.conbuildmat.2015.10.026
Ismail, I., Provis, J.L., Deventer, J.S.J. Van, Hamdan., S., 2011. The effect of water content on
compressive strength of geopolymer mortars, in: The AES-ATEMA’2011 International
Conference on Advances and Trends in Engineering Materials and Their Applications.
Jain, M.K., Pal, S., 1998. Utilisation of industrial slag in making high performance concrete
composites. Indian Concr. J. 72, 307–315.
Johnson, G., 2007. Geopolymer concrete and method of preparation and casting.
Joseph, B., Mathew, G., 2012. Influence of aggregate content on the behavior of fly ash based
geopolymer concrete. Sci. Iran. https://doi.org/10.1016/j.scient.2012.07.006
Junaid, M.T., Khennane, A., Kayali, O., Sadaoui, A., Picard, D., Fafard, M., 2014. Aspects of the
deformational behaviour of alkali activated fly ash concrete at elevated temperatures. Cem.
Concr. Res. 60, 24–29. https://doi.org/10.1016/j.cemconres.2014.01.026
Kabir, S.M.A., Alengaram, U.J., Jumaat, M.Z., Sharmin, A., Islam, A., 2015. Influence of
molarity and chemical composition on the development of compressive strength in POFA
based geopolymer mortar. Adv. Mater. Sci. Eng. 2015, 1–15.
https://doi.org/10.1155/2015/647071
Kamath, A., Khan, M.S., 2016. Green concrete. Imp. J. Interdiscip. Res. 3.
Kamseu, E., Ponzoni, C., Tippayasam, C., Taurino, R., Chaysuwan, D., Sglavo, V.M.,
Thavorniti, P., Leonelli, C., 2016. Self-compacting geopolymer concretes: Effects of
addition of aluminosilicate-rich fines. J. Build. Eng.
https://doi.org/10.1016/j.jobe.2016.01.004
Karbhari, V.M., 2013. Non-Destructive Evaluation (Nde) Of Polymer Matrix Composites.
Elsevier Science.
Kavitha, O.R., Shanthi, V.M., Arulraj, G.P., Sivakumar, V.R., 2016. Microstructural studies on
eco-friendly and durable Self-compacting concrete blended with metakaolin. Appl. Clay
Sci. https://doi.org/10.1016/j.clay.2016.02.011
Kelham, S., 1996. The influence of high early-strength (HES) mineralized clinker on the strength
of development of blended cements containing fly ash, slag, or ground limestone. Fuel
Energy Abstr.
Khale, D., Chaudhary, R., 2007. Mechanism of geopolymerization and factors influencing its
development: A review. J. Mater. Sci. 42, 729–746. https://doi.org/10.1007/s10853-006-
0401-4

74
Journal Pre-proof

Khandelwal, M., Ranjith, P.G., Pan, Z., Sanjayan, J.G., 2013. Effect of strain rate on strength
properties of low-calcium fly-ash-based geopolymer mortar under dry condition. Arab. J.
Geosci. 6, 2383–2389. https://doi.org/10.1007/s12517-011-0507-0
Khayat, K., Manai, K., Trudel, A., 1997. In situ mechanical properties of wall elements cast
using self-consolidating concrete. ACI Mater. J. 94, 491–500.
Khayat, K.H., Guizani, Z., 1997. Use of viscosity-modifying admixture to enhance stability of
fluid concrete. ACI Mater. J. 94, 332–340.
Kirupa, J.A.D., Sakthieswaran, N., 2015. Possible materials for producing Geopolymer concrete
and its performance with and without Fibre addition-A State of the art review. Int. J. Civ.
Struct. Eng. 5, 296.
Klabunde, K.J., Richards, R., 2009. Nanoscale Materials in Chemistry, Nanoscale Materials in
Chemistry. Wiley. https://doi.org/10.1002/9780470523674
Komnitsas, K., Zaharaki, D., 2007. Geopolymerisation: A review and prospects for the minerals
industry. Miner. Eng. 20, 1261–1277. https://doi.org/10.1016/j.mineng.2007.07.011
Kong, D.., Sanjayan, J.G., Sagoe-Crentsil., K., 2007. Comparative performance of geopolymers
made with metakaolin and fly ash after exposure to elevated temperatures. Cem. Concr.
Res. 37, 1583–1589.
Kong, D.L.Y., Sanjayan, J.G., 2010. Effect of elevated temperatures on geopolymer paste,
mortar and concrete. Cem. Concr. Res. 40, 334–339.
https://doi.org/10.1016/j.cemconres.2009.10.017
Kou, S.-C., Poon, C.-S., Wan, H.-W., 2012. Properties of concrete prepared with low-grade
recycled aggregates. Constr. Build. Mater. 36, 881–889.
https://doi.org/10.1016/j.conbuildmat.2012.06.060
Kovacik, P., Macák, M., Duscay, L., Halcinova, M., Jancich, M., 2011. Effect of ash-fly ash
mixture application on soil fertility. J. Elem. 16.
Kovler, K., Roussel, N., 2011. Properties of fresh and hardened concrete. Cem. Concr. Res.
https://doi.org/10.1016/j.cemconres.2011.03.009
KThu, M., Murthy, D.R., 2015. An Experimental Investigation on the Mechanical Properties of
Geopolymer Concrete Partially Replaced with Recycled Coarse Aggregates. Int. J. Sci. Eng.
Res. 6.
Kumar, A., Kumar, S., 2013. Development of paving blocks from synergistic use of red mud and
fly ash using geopolymerization. Constr. Build. Mater. 38, 865–871.
https://doi.org/10.1016/j.conbuildmat.2012.09.013
Kumaravel, S., 2014. Development of various curing effect of nominal strength Geopolymer
concrete. J. Eng. Sci. Technol. Rev. 7, 116–119. https://doi.org/10.25103/jestr.071.19
Lakshmi, R., Nagan, S., 2011. Utilization of waste e plastic particles in cementitious mixtures. J.
Struct. Eng.
Law, D.W., Adam, A.A., Molyneaux, T.K., Patnaikuni, I., Wardhono, A., 2015. Long term
durability properties of class F fly ash geopolymer concrete. Mater. Struct. 48, 721–731.
https://doi.org/10.1617/s11527-014-0268-9
Lee, N.K., Kim, E.M., Lee, H.K., 2016. Mechanical properties and setting characteristics of
geopolymer mortar using styrene-butadiene (SB) latex. Constr. Build. Mater. 113, 264–272.
https://doi.org/10.1016/j.conbuildmat.2016.03.055
Leong, H.Y., Ong, D.E.L., Sanjayan, J.G., Nazari, A., 2016. Suitability of Sarawak and
Gladstone fly ash to produce geopolymers: A physical, chemical, mechanical, mineralogical
and microstructural analysis. Ceram. Int. 42, 9613–9620.

75
Journal Pre-proof

https://doi.org/10.1016/j.ceramint.2016.03.046
Li, G., Zhao, X., 2003. Properties of concrete incorporating fly ash and ground granulated blast-
furnace slag. Cem. Concr. Compos. 25, 293–299. https://doi.org/10.1016/S0958-
9465(02)00058-6
Li, H., Liu, G., Cao, Y., 2014. Content and distribution of trace elements and polycyclic aromatic
hydrocarbons in fly ash from a coal-fired CHP plant. Aerosol Air Qual. Res.
https://doi.org/10.4209/aaqr.2013.06.0216
Li, W., Xu, J., 2009. Mechanical properties of basalt fiber reinforced geopolymeric concrete
under impact loading. Mater. Sci. Eng. A 505, 178–186.
https://doi.org/10.1016/j.msea.2008.11.063
Liang, K., Jin, S., Chen, H., Ren, J., Shen, W., Wei, S., 2019. Parametric optimization of packed
bed for activated coal Fly ash waste heat recovery using CFD techniques. Chinese J. Chem.
Eng. https://doi.org/10.1016/j.cjche.2019.06.004
Liu, M.Y.J., Alengaram, U.J., Santhanam, M., Jumaat, M.Z., Mo, K.H., 2016. Microstructural
investigations of palm oil fuel ash and fly ash based binders in lightweight aggregate
foamed geopolymer concrete. Constr. Build. Mater. 120, 112–122.
https://doi.org/10.1016/j.conbuildmat.2016.05.076
Liu, W., Yang, J., Xiao, B., 2009. Review on treatment and utilization of bauxite residues in
China. Int. J. Miner. Process. 93, 220–231. https://doi.org/10.1016/j.minpro.2009.08.005
Lloyd, R.R., Provis, J.L., van Deventer, J.S.J., 2012. Acid resistance of inorganic polymer
binders. 1. Corrosion rate. Mater. Struct. 45, 1–14. https://doi.org/10.1617/s11527-011-
9744-7
M. Greim, W. Kusterle, 2004. Rheological Measurement of Building Materials (Regensburg,
Germany). Appl. Rheol. 14, 148–150. https://doi.org/10.3933/ApplRheol-14-148
Ma, Y., Hu, J., Ye, G., 2013. The pore structure and permeability of alkali activated fly ash. Fuel
104, 771–780. https://doi.org/10.1016/j.fuel.2012.05.034
Madheswaran, C., Gnanasundar, G., Gopalakrishnan, N., 2013. Effect of molarity in geopolymer
concrete. Int. J. Civ. Struct. Eng. 4, 106–115. https://doi.org/10.6088/ijcser.20130402001
Malathy, D.., 2009. Fresh And Hardened Properties of Geo Polymer Concrete And Mortar. Civ.
Eng. Kongu engg. Coll. Perundurai, India.
Mander, J.B., Priestley, M.J.N., Park, R., 1988. Theoretical Stress‐Strain Model for Confined
Concrete. J. Struct. Eng. 114, 1804–1826. https://doi.org/10.1061/(ASCE)0733-
9445(1988)114:8(1804)
MATSAGAR, V. (Ed. ., 2015. Advances in Structural Engineering, Advances in Structural
Engineering. Springer India, New Delhi.
Mazaheripour, H., Ghanbarpour, S., Mirmoradi, S.H., Hosseinpour, I., 2011. The effect of
polypropylene fibers on the properties of fresh and hardened lightweight self-compacting
concrete. Constr. Build. Mater. https://doi.org/10.1016/j.conbuildmat.2010.06.018
Mazloom, M., Ramezanianpour, A.A., Brooks, J.J., 2004. Effect of silica fume on mechanical
properties of high-strength concrete. Cem. Concr. Compos. 26, 347–357.
https://doi.org/10.1016/S0958-9465(03)00017-9
Mehta, P.., 1977. Properties of blended cements made from rice husk ash. J. Proc.
Memon, F.A., Nuruddin, M.F., Demie, S., Shafiq, N., 2012. Effect of superplasticizer and extra
water on workability and compressive strength of self-compacting geopolymer concrete.
Res. J. Appl. Sci. Eng. Technol. 4, 407–414.
Memon, F.A., Nuruddin, M.F., Shafiq, N., 2013. Effect of silica fume on the fresh and hardened

76
Journal Pre-proof

properties of fly ash-based self-compacting geopolymer concrete. Int. J. Miner. Metall.


Mater. 20, 205–213. https://doi.org/10.1007/s12613-013-0714-7
Mo, K.H., Alengaram, U.J., Jumaat, M.Z., 2015a. Experimental Investigation on the Properties
of Lightweight Concrete Containing Waste Oil Palm Shell Aggregate. Procedia Eng. 125,
587–593. https://doi.org/10.1016/j.proeng.2015.11.065
Mo, K.H., Alengaram, U.J., Jumaat, M.Z., Liu, M.Y.J., Lim, J., 2016. Assessing some durability
properties of sustainable lightweight oil palm shell concrete incorporating slag and
manufactured sand. J. Clean. Prod. 112, 763–770.
https://doi.org/10.1016/j.jclepro.2015.06.122
Mo, K.H., Alengaram, U.J., Visintin, P., Goh, S.H., Jumaat, M.Z., 2015b. Influence of
lightweight aggregate on the bond properties of concrete with various strength grades.
Constr. Build. Mater. 84, 377–386. https://doi.org/10.1016/j.conbuildmat.2015.03.040
Moghadam, H.A., Khoshbin, O.A., 2012. Effect of Water- Cement Ratio (w/c) on Mechanical
Properties of Self-Compacting Concrete (Case Study).
Muthupriya, P., Manjunath, N., Keerdhana, B., 2014. Strength study on fiber reinforced self-
compacting concrete with fly ash and GGBFS. Int. J. Adv. Struct. Geotech. Eng. 3.
Naik, T.R., 2008. Sustainability of concrete construction. Pract. Period. Struct. Des. Constr.
https://doi.org/10.1061/(ASCE)1084-0680(2008)13:2(98)
Naji, G.A., Rashid, S.A., Aziz, F.N.A., Salleh, M.A.M., 2010. Contribution of rice husk ash to
the properties of mortar and concrete: a review. J. Am. Sci. 6, 157–165.
Nath, P., Sarker, P.K., 2015. Use of OPC to improve setting and early strength properties of low
calcium fly ash geopolymer concrete cured at room temperature. Cem. Concr. Compos. 55,
205–214. https://doi.org/10.1016/j.cemconcomp.2014.08.008
Nath, P., Sarker, P.K., 2012. Geopolymer concrete for ambient curing condition, in: In
Australasian Structural Engineering Conference 2012: The Past, Present and Future of
Structural Engineering. Engineers Australia, p. 225.
Nath, S.K., Kumar, S., 2013. Influence of iron making slags on strength and microstructure of fly
ash geopolymer. Constr. Build. Mater. 38, 924–930.
https://doi.org/10.1016/j.conbuildmat.2012.09.070
Nazari, A., Bagheri, A., Riahi, S., 2011. Properties of geopolymer with seeded fly ash and rice
husk bark ash. Mater. Sci. Eng. A 528, 7395–7401.
https://doi.org/10.1016/j.msea.2011.06.027
Nematollahi, B., Qiu, J., Yang, E.-H., Sanjayan, J., 2017. Micromechanics constitutive
modelling and optimization of strain hardening geopolymer composite. Ceram. Int. 43,
5999–6007. https://doi.org/10.1016/j.ceramint.2017.01.138
Nematollahi, B., Sanjayan, J., 2014. Effect of different superplasticizers and activator
combinations on workability and strength of fly ash based geopolymer. Mater. Des. 57,
667–672. https://doi.org/10.1016/j.matdes.2014.01.064
Neville, A.M., 1995. Properties of concrete. Prentice Hall.
Neville, A.M. and J.., 1987. Brooks. Concrete technology.
Noushini, A., Aslani, F., Castel, A., Gilbert, R.I., Uy, B., Foster, S., 2016. Compressive stress-
strain model for low-calcium fly ash-based geopolymer and heat-cured Portland cement
concrete. Cem. Concr. Compos. https://doi.org/10.1016/j.cemconcomp.2016.07.004
Nuaklong, P., Sata, V., Chindaprasirt, P., 2016. Influence of recycled aggregate on fly ash
geopolymer concrete properties. J. Clean. Prod.
https://doi.org/10.1016/j.jclepro.2015.10.109

77
Journal Pre-proof

Olsson, J., Jernqvist, Å., Aly, G., 1997. Thermophysical properties of aqueous NaOH−H2O
solutions at high concentrations. Int. J. Thermophys. 18, 779–793.
https://doi.org/10.1007/BF02575133
Palacios, M., Puertas, F., 2004. Stability of superplasticizer and shrinkage-reducing admixtures
in high basic media. Mater. Construcción 54, 65–86.
PARAMGURU, R.K., RATH, P.C., MISRA, V.N., 2004. TRENDS IN RED MUD
UTILIZATION – A REVIEW. Miner. Process. Extr. Metall. Rev. 26, 1–29.
https://doi.org/10.1080/08827500490477603
Patil, A.A., Chore, H., Dodeb, P., 2014. Effect of curing condition on strength of geopolymer
concrete. Adv. Concr. Constr. 2, 29–37.
Pauling, L., 1988. General chemistry. Courier Corporation.
Phoo-ngernkham, T., Chindaprasirt, P., Sata, V., Hanjitsuwan, S., Hatanaka, S., 2014. The effect
of adding nano-SiO2 and nano-Al2O3 on properties of high calcium fly ash geopolymer
cured at ambient temperature. Mater. Des. 55, 58–65.
https://doi.org/10.1016/j.matdes.2013.09.049
Poon, C.S., Kou, S.C., Lam, L., 2006. Compressive strength, chloride diffusivity and pore
structure of high performance metakaolin and silica fume concrete. Constr. Build. Mater.
20, 858–865. https://doi.org/10.1016/j.conbuildmat.2005.07.001
Posi, P., Teerachanwit, C., Tanutong, C., Limkamoltip, S., Lertnimoolchai, S., Sata, V.,
Chindaprasirt, P., 2013. Lightweight geopolymer concrete containing aggregate from
recycle lightweight block. Mater. Des. https://doi.org/10.1016/j.matdes.2013.06.001
Poudenx, P., 2008. The effect of transportation policies on energy consumption and greenhouse
gas emission from urban passenger transportation. Transp. Res. Part A Policy Pract. 42,
901–909. https://doi.org/10.1016/j.tra.2008.01.013
Pouhet, R., Cyr, M., 2016. Formulation and performance of flash metakaolin geopolymer
concretes. Constr. Build. Mater. 120, 150–160.
https://doi.org/10.1016/j.conbuildmat.2016.05.061
Prabhu, R.., Anuradha, R., S. Vivek, 2016. Experimental Research on Triple Blended Self-
Compacting Geo Polymer Concrete. Asian J. Eng. Appl. Technol. 5, 15–21.
Prediction of Long-Term Corrosion Resistance of Plain and Blended Cement concretes, 1993. .
ACI Mater. J. https://doi.org/10.14359/4430
Provis, J.L., Deventer, J.S.J. Van, 2009. Geopolymers: structures, processing, properties and
industrial applications. Elsevier.
Provis, J.L., Lukey, G.C., van Deventer, J.S.J., 2005. Do Geopolymers Actually Contain
Nanocrystalline Zeolites? A Reexamination of Existing Results. Chem. Mater. 17, 3075–
3085. https://doi.org/10.1021/cm050230i
Provis, J.L., Myers, R.J., White, C.E., Rose, V., Van Deventer, J.S.J., 2012. X-ray
microtomography shows pore structure and tortuosity in alkali-activated binders. Cem.
Concr. Res. https://doi.org/10.1016/j.cemconres.2012.03.004
Prud’homme, E., Joussein, E., Rossignol, S., 2015. Use of silicon carbide sludge to form porous
alkali-activated materials for insulating application. Eur. Phys. J. Spec. Top. 224, 1725–
1735. https://doi.org/10.1140/epjst/e2015-02494-7
Puertas, F., Palomo, A., Fernández-Jiménez, A., Izquierdo, J.D., Granizo, M.L., 2003. Effect of
superplasticisers on the behaviour and properties of alkaline cements. Adv. Cem. Res. 15,
23–28. https://doi.org/10.1680/adcr.2003.15.1.23
Raheem, M.., Otuose, H.S., Abdulhafiz, U., 2013. Properties of Rice Husk Ash Stabilized

78
Journal Pre-proof

Laterite Roof Tiles. Leonardo Electron. J. Pract. Technol. 23, 41–50.


Reddy, E.B.S.S., 2015. Geo Polymer Concrete with the Replacement of Granite Aggregate as
Fine Aggregate. Int. J. Sci. Res. 4, 19–24.
Rickard, W.., Gluth, G.J., Pistol, K., 2016. In-situ thermo-mechanical testing of fly ash
geopolymer concretes made with quartz and expanded clay aggregates. Cem. Concr. Res.
80, 33–43.
Rickard, W., Vickers, L., Riessen., A. Van, 2014. Fire-Resistant Geopolymers. Springer.
Ridtirud, C., Chindaprasirt, P., Pimraksa, K., 2011. Factors affecting the shrinkage of fly ash
geopolymers. Int. J. Miner. Metall. Mater. 18, 100–104. https://doi.org/10.1007/s12613-
011-0407-z
RUBEL, A., ANDREWS, R., GONZALEZ, R., GROPPO, J., ROBL, T., 2005. Adsorption of
Hg and NO on coal by-products. Fuel 84, 911–916.
https://doi.org/10.1016/j.fuel.2005.01.006
Ryu, G.S., Lee, Y.B., Koh, K.T., Chung, Y.S., 2013. The mechanical properties of fly ash-based
geopolymer concrete with alkaline activators. Constr. Build. Mater. 47, 409–418.
https://doi.org/10.1016/j.conbuildmat.2013.05.069
Sahithi, K., Priyanka., G.S., 2015. The Effect of Addition of Limestone powder and Granulated
Blast Slag in Concrete. Int. J. Comput. Eng. Res. Trends 2.
Sakai, E., 1997. The fluidity of cement paste with various types of inorganic powders, in: Proc.
of the 10th Int’l Cong. on the Chemistry of Cement. Gothenburg, Sweden.
Sakulich, A.R., 2011. Reinforced geopolymer composites for enhanced material greenness and
durability,. Sustain. Cities Soc. 1, 195–210.
Salih, M.A., Abang Ali, A.A., Farzadnia, N., 2014. Characterization of mechanical and
microstructural properties of palm oil fuel ash geopolymer cement paste. Constr. Build.
Mater. 65, 592–603. https://doi.org/10.1016/j.conbuildmat.2014.05.031
San Nicolas, R., Cyr, M., Escadeillas, G., 2013. Characteristics and applications of flash
metakaolins. Appl. Clay Sci. 83–84, 253–262. https://doi.org/10.1016/j.clay.2013.08.036
Sanjayan, J.G., Nazari, A., Pouraliakbar, H., 2015. FEA modelling of fracture toughness of steel
fibre-reinforced geopolymer composites. Mater. Des. 76, 215–222.
https://doi.org/10.1016/j.matdes.2015.03.029
Saravanan, M., Sivaraja, M., 2016. Mechanical Behavior of Concrete Modified by Replacement
of Cement by Rice Husk Ash. Brazilian Arch. Biol. Technol. 59(SPE2).
Saraya, M.E.-S.I., 2014. Study physico-chemical properties of blended cements containing fixed
amount of silica fume, blast furnace slag, basalt and limestone, a comparative study. Constr.
Build. Mater. 72, 104–112. https://doi.org/10.1016/j.conbuildmat.2014.08.071
Sashidhar, C., J. Guru Jawahar, C. Neelima, D. Pavan Kumar, 2015. Fresh and strength
properties of self compacting geopolymer concrete using manufactured sand. Int. J.
ChemTech Res. 8, 183–190.
Sashidhar, C., J., J., Guru, C., Neelima, D, K., 2016. Preliminary studies on self-compacting
geopolymer concrete using manufactured sand. Asian J. Civ. Eng. 17, 277–288.
Sata, V., Sathonsaowaphak, A., Chindaprasirt, P., 2012. Resistance of lignite bottom ash
geopolymer mortar to sulfate and sulfuric acid attack. Cem. Concr. Compos.
https://doi.org/10.1016/j.cemconcomp.2012.01.010
Sata, V., Wongsa, A., Chindaprasirt, P., 2013. Properties of pervious geopolymer concrete using
recycled aggregates. Constr. Build. Mater.
https://doi.org/10.1016/j.conbuildmat.2012.12.046

79
Journal Pre-proof

Satpute Manesh, B., Madhukar, R.W., Subhash., V.P., 2012. Effect of duration and temperature
of curing on compressive strength of geopolymer concrete. Int. J. Eng. Innov. Technol. 1.
Shaikh, F.U.A., 2016. Mechanical and durability properties of fly ash geopolymer concrete
containing recycled coarse aggregates. Int. J. Sustain. Built Environ. 5, 277–287.
https://doi.org/10.1016/j.ijsbe.2016.05.009
Shaikh, F.U.A., 2014. Effects of alkali solutions on corrosion durability of geopolymer concrete.
Adv. Concr. Constr. 2, 109–123. https://doi.org/10.12989/acc.2014.2.2.109
Shaikh, F.U.A., Odoh, H., Than, A.B., 2015. Effect of nano silica on properties of concretes
containing recycled coarse aggregates. Proc. Inst. Civ. Eng. - Constr. Mater. 168, 68–76.
https://doi.org/10.1680/coma.14.00009
Shalini, A., Gurunarayanan, G., Kumar, R.., Prakash, V.., Sakthivel, S., 2016. Performance of
Rice Husk Ash in Geopolymer Concrete. Int. J. Innov. Res. Sci. Technol. 2, 73–77.
Sharafeddin, F., Alavi, A., Talei, Z., 2013. Flexural strength of glass and polyethylene fiber
combined with three different composites. J. Dent. (Shiraz, Iran) 14, 13–9.
Shi, C., Jiménez, A.F., Palomo, A., 2011. New cements for the 21st century: The pursuit of an
alternative to Portland cement. Cem. Concr. Res.
https://doi.org/10.1016/j.cemconres.2011.03.016
Shi, C., Wu, Z., Xiao, J., Wang, D., Huang, Z., Fang, Z., 2015. A review on ultra high
performance concrete: Part I. Raw materials and mixture design. Constr. Build. Mater. 101,
741–751. https://doi.org/10.1016/j.conbuildmat.2015.10.088
Shi, X.S., Wang, Q.Y., Zhao, X.L., Collins, F., 2012. Discussion on Properties and
Microstructure of Geopolymer Concrete Containing Fly Ash and Recycled Aggregate. Adv.
Mater. Res. 450–451, 1577–1583. https://doi.org/10.4028/www.scientific.net/AMR.450-
451.1577
Shrivastava, S., Shrivastava., R., 2015. Convergence of the sequence of ishikawa iteration
process with errors for fixed points. Int. J. Adv. Technol. Eng. Sci. 3.
Shukla, J.P., Mondal, D.P., Jain, P.K., Shukla, A., 2009. New Technologies for Rural
Development Having Potential of Commercialisation. Allied Publ.
Si, C., Ma, Y., Lin, C., 2013. Red mud as a carbon sink: Variability, affecting factors and
environmental significance. J. Hazard. Mater. 244–245, 54–59.
https://doi.org/10.1016/j.jhazmat.2012.11.024
Siddique, R., Iqbal Khan, M., 2011. Supplementary Cementing Materials.
Singh, B., Ishwarya, G., Gupta, M., Bhattacharyya, S.K., 2015. Geopolymer concrete: A review
of some recent developments. Constr. Build. Mater. 85, 78–90.
https://doi.org/10.1016/j.conbuildmat.2015.03.036
Singh, M., Siddique, R., 2015. Properties of concrete containing high volumes of coal bottom
ash as fine aggregate. J. Clean. Prod. 91, 269–278.
https://doi.org/10.1016/j.jclepro.2014.12.026
Sivaraja, M., Kandasamy, Velmani, N., Pillai, M.S., 2010. Study on durability of natural fibre
concrete composites using mechanical strength and microstructural properties. Bull. Mater.
Sci. 33, 719–729. https://doi.org/10.1007/s12034-011-0149-6
Sobolev, K., Gutiérrez, M.F., 2005. How nanotechnology can change the concrete world. Am.
Ceram. Soc. Bull. 84, 14–18. https://doi.org/10.1002/9780470588260.ch16
Soltaninaveh, K., 2008. The Properties of Geopolymer Concrete Incorporating Red Sand as Fine
Aggregate. Test.
Somna, K., Jaturapitakkul, C., Kajitvichyanukul, P., Chindaprasirt, P., 2011. NaOH-activated

80
Journal Pre-proof

ground fly ash geopolymer cured at ambient temperature. Fuel 90, 2118–2124.
https://doi.org/10.1016/j.fuel.2011.01.018
Soroka, I., Jaegermann, C.H., Bentur, A., 1978. Short-term steam-curing and concrete later-age
strength. Matériaux Constr. 11, 93–96. https://doi.org/10.1007/BF02478955
Soutsos, M., Boyle, A.P., Vinai, R., Hadjierakleous, A., Barnett, S.J., 2016. Factors influencing
the compressive strength of fly ash based geopolymers. Constr. Build. Mater. 110, 355–368.
https://doi.org/10.1016/j.conbuildmat.2015.11.045
Spanlang, A., Wukovits, W., Weiss, B., 2016. Development of a blast furnace model with
thermodynamic process depiction by means of the rist operating diagram. Chem. Eng.
Trans. https://doi.org/10.3303/CET1652163
Sreenivasulu, C., A.R., Jawahar, J.G., 2015. Mechanical properties of geopolymer concrete using
granite slurry as sand replacement. Int. J. Adv. Eng. Technol. 8, 83.
Su, H., Xu, J., Ren, W., 2016. Mechanical properties of geopolymer concrete exposed to
dynamic compression under elevated temperatures. Ceram. Int. 42, 3888–3898.
https://doi.org/10.1016/j.ceramint.2015.11.055
Subang Jaya, S., NURUDDIN, M.F., KHAN, S., SHAFIQ, N., AYUB, T., 2013. Effect of
sodium hydroxide concentration on fresh properties and compressive strength of self-
compacting geopolymer concrete. J. Eng. Sci. Technol. 8, 44–56.
Sumajouw, D.M.J., Hardjito, D., Wallah, S.E., Rangan, B. V., 2004. Geopolymer concrete for a
sustainable future. Present. Green Process. Conf. Fremantle, WA.
Suresh, D., Nagaraju, K., 2015. Ground Granulated Blast Slag (GGBS) In Concrete-A Review.
IOSR J. Mech. Civ. Eng. 12, 76–82. https://doi.org/10.9790/1684-12467682
Talakokula, V., Vaibhav, Bhalla, S., 2016. Non-destructive Strength Evaluation of Fly Ash
Based Geopolymer Concrete Using Piezo Sensors, in: Procedia Engineering. Elsevier, pp.
1029–1035. https://doi.org/10.1016/j.proeng.2016.04.133
Temuujin, J., Minjigmaa, A., Lee, M., Chen-Tan, N., van Riessen, A., 2011. Characterisation of
class F fly ash geopolymer pastes immersed in acid and alkaline solutions. Cem. Concr.
Compos. 33, 1086–1091. https://doi.org/10.1016/j.cemconcomp.2011.08.008
Thampi, T., Sreevidya, V., Venkatasubramani, R., 2014. Strength studies on geopolymer mortar
for ferro-geopolymer water tank. Int. J. Adv. Struct. Geotech. Eng. 3, 102–105.
Thokchom, S., Ghosh, P., Ghosh, S., 2009. Effect of water absorption, porosity and sorptivity on
durability of geopolymer mortars. ARPN J. Eng. Appl. Sci. 4, 28–32.
Thomas, R.J., Peethamparan, S., 2015. Alkali-activated concrete: Engineering properties and
stress–strain behavior. Constr. Build. Mater. 93, 49–56.
https://doi.org/10.1016/j.conbuildmat.2015.04.039
Tolêdo Filho, R.D., Ghavami, K., England, G.L., Scrivener, K., 2003. Development of vegetable
fibre–mortar composites of improved durability. Cem. Concr. Compos. 25, 185–196.
https://doi.org/10.1016/S0958-9465(02)00018-5
Topark-Ngarm, P., Chindaprasirt, P., Sata, V., 2014. Setting Time, Strength, and Bond of High-
Calcium Fly Ash Geopolymer Concrete. J. Mater. Civ. Eng.
https://doi.org/10.1061/(asce)mt.1943-5533.0001157
Triantafillou, T., 2016. Textile fibre composites in civil engineering.
Ukwattage, N.L., Ranjith, P.G., Bouazza, M., 2013. The use of coal combustion fly ash as a soil
amendment in agricultural lands (with comments on its potential to improve food security
and sequester carbon). Fuel 109, 400–408. https://doi.org/10.1016/j.fuel.2013.02.016
Usha, S., Nair, D.G., Vishnudas, S., 2016. Feasibility Study of Geopolymer Binder from

81
Journal Pre-proof

Terracotta Roof Tile Waste. Procedia Technol. 25, 186–193.


https://doi.org/10.1016/j.protcy.2016.08.096
Ushaa, T., Anuradha, R., Venkatasubramani, G., 2015. Performance of self-compacting
geopolymer concrete containing different mineral admixtures. Indian J. Eng. Mater. Sci. 22.
Vaidya, S., Allouche, E.N., 2011. Strain sensing of carbon fiber reinforced geopolymer concrete.
Mater. Struct. 44, 1467–1475. https://doi.org/10.1617/s11527-011-9711-3
van Jaarsveld, J., Deventer, J. Van, 1999. Effect of the alkali metal activator on the properties of
fly ash-based geopolymers. Ind. Eng. Chem. Res. 38, 3932–3941.
Venkatanarayanan, H.K., Rangaraju, P.R., 2014. Evaluation of Sulfate Resistance of Portland
Cement Mortars Containing Low-Carbon Rice Husk Ash. J. Mater. Civ. Eng. 26, 582–592.
https://doi.org/10.1061/(ASCE)MT.1943-5533.0000868
Vijai, K., Kumutha, R., Vishnuram, B.G., 2010. Effect of types of curing on strength of
geopolymer concrete. Int. J. Phys. Sci.
Wardhono, A., Gunasekara, C., Law, D.W., Setunge, S., 2017. Comparison of long term
performance between alkali activated slag and fly ash geopolymer concretes. Constr. Build.
Mater. 143, 272–279. https://doi.org/10.1016/j.conbuildmat.2017.03.153
Wolsiefer, J.T., 1991. Silica fume concrete: a solution to steel reinforcement corrosion in
concrete. Int. Conf. Durab. Concr.
Wongpa, J., Kiattikomol, K., Jaturapitakkul, C., Chindaprasirt, P., 2010. Compressive strength,
modulus of elasticity, and water permeability of inorganic polymer concrete. Mater. Des.
https://doi.org/10.1016/j.matdes.2010.05.012
Yan, L., Kasal, B., Huang, L., 2016. A review of recent research on the use of cellulosic fibres,
their fibre fabric reinforced cementitious, geo-polymer and polymer composites in civil
engineering. Compos. Part B Eng. 92, 94–132.
https://doi.org/10.1016/j.compositesb.2016.02.002
Yan, S., Sagoe-Crentsil, K., 2012. Properties of wastepaper sludge in geopolymer mortars for
masonry applications. J. Environ. Manage. 112, 27–32.
Yang, K.H., Song, J.K., Song, K. Il, 2013. Assessment of CO 2 reduction of alkali-activated
concrete. J. Clean. Prod. 39, 265–272. https://doi.org/10.1016/j.jclepro.2012.08.001
Ye, N., Yang, J., Ke, X., Zhu, J., Li, Y., Xiang, C., Wang, H., Li, L., Xiao, B., 2014. Synthesis
and Characterization of Geopolymer from Bayer Red Mud with Thermal Pretreatment. J.
Am. Ceram. Soc. 97, 1652–1660. https://doi.org/10.1111/jace.12840
Ye, N., Yang, J., Liang, S., Hu, Y., Hu, J., Xiao, B., Huang, Q., 2016. Synthesis and strength
optimization of one-part geopolymer based on red mud. Constr. Build. Mater. 111, 317–
325. https://doi.org/10.1016/j.conbuildmat.2016.02.099
Yildiz, E., 2004. Phosphate removal from water by fly ash using crossflow microfiltration. Sep.
Purif. Technol. 35, 241–252. https://doi.org/10.1016/S1383-5866(03)00145-X
Yusuf, M.O., Megat Johari, M.A., Ahmad, Z.A., Maslehuddin, M., 2015. Impacts of silica
modulus on the early strength of alkaline activated ground slag/ultrafine palm oil fuel ash
based concrete. Mater. Struct. 48, 733–741. https://doi.org/10.1617/s11527-014-0318-3
Zaharieva, R., Buyle-Bodin, F., Skoczylas, F., Wirquin, E., 2003. Assessment of the surface
permeation properties of recycled aggregate concrete. Cem. Concr. Compos. 25, 223–232.
https://doi.org/10.1016/S0958-9465(02)00010-0
Zain, M.F.M., Islam, M.N., Mahmud, F., Jamil, M., 2011. Production of rice husk ash for use in
concrete as a supplementary cementitious material. Constr. Build. Mater.
https://doi.org/10.1016/j.conbuildmat.2010.07.003

82
Journal Pre-proof

Zhang, J.J., Bentley, L.R., 2003. Pore geometry and elastic moduli in sandstones.
Zhang, W.X., Wang, C.B., Lien, H.L., 1998. Treatment of chlorinated organic contaminants with
nanoscale bimetallic particles. Catal. Today 40, 387–395. https://doi.org/10.1016/S0920-
5861(98)00067-4
Zhang, Y.J., Wang, Y.C., Xu, D.L., Li, S., 2010. Mechanical performance and hydration
mechanism of geopolymer composite reinforced by resin. Mater. Sci. Eng. A 527, 6574–
6580. https://doi.org/10.1016/j.msea.2010.06.069
Zhang, Z., Provis, J.L., Zou, J., Reid, A., Wang, H., 2016. Toward an indexing approach to
evaluate fly ashes for geopolymer manufacture. Cem. Concr. Res. 85, 163–173.
https://doi.org/10.1016/j.cemconres.2016.04.007
Zhuang, X.Y., Chen, L., Komarneni, S., Zhou, C.H., Tong, D.S., Yang, H.M., Yu, W.H., Wang,
H., 2016a. Fly ash-based geopolymer: Clean production, properties and applications. J.
Clean. Prod. https://doi.org/10.1016/j.jclepro.2016.03.019
Zhuang, X.Y., Chen, L., Komarneni, S., Zhou, C.H., Tong, D.S., Yang, H.M., Yu, W.H., Wang,
H., 2016b. Fly ash-based geopolymer: clean production, properties and applications. J.
Clean. Prod. 125, 253–267. https://doi.org/10.1016/J.JCLEPRO.2016.03.019

83
Journal Pre-proof

Conflict of Interest and Authorship Conformation Form

Please check the following as appropriate:

 All authors have participated in (a) conception and design, or analysis and
interpretation of the data; (b) drafting the article or revising it critically for
important intellectual content; and (c) approval of the final version.

 This manuscript has not been submitted to, nor is under review at, another
journal or other publishing venue.

 The authors have no affiliation with any organization with a direct or indirect
financial interest in the subject matter discussed in the manuscript

 The following authors have affiliations with organizations with direct or


indirect financial interest in the subject matter discussed in the manuscript:

Author’s name Affiliation Signature


Y. H. Mugahed Amran Prince Sattam Bin Abdulaziz University, Saudi Arabia
Amran University, Yemen and
Rayed Alyousef Prince Sattam Bin Abdulaziz University, Saudi Arabia
Hisham Alabduljabbar Prince Sattam Bin Abdulaziz University, Saudi Arabia
Mohamed El-Zeadani Universiti Putra Malaysia
Journal Pre-proof

Highlights

 Geopolymer concrete has a distinguished property of being reduced the use of vibration and cement.
 The main function of developing geopolymer concrete is to reduce the emissions of CO2 in the
atmosphere.
 The rate of strength development of geopolymer concrete is mainly relied on the source of materials,
alkaline activators, and curing conditions.
 Geopolymer concrete has emerged as a new engineering material, contributing towards
environmentally sustainable construction and building products industry.

You might also like