Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/333798778

Geopolymer concrete at ambient and elevated temperatures: Recent


developments and challenges

Preprint · July 2019


DOI: 10.35453/NEDJR-STMECH-2019-0007

CITATIONS READS

0 435

2 authors, including:

Professor Zhong Tao


Western Sydney University
176 PUBLICATIONS   3,407 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Geopolymer concrete and its applications View project

Development of sustainable houses View project

All content following this page was uploaded by Professor Zhong Tao on 15 June 2019.

The user has requested enhancement of the downloaded file.


This is an electronic version of an article published in NED University Journal of Research,
which is available online at:
https://www.neduet.edu.pk/NED-Journal/2019/fullpage.html
Tao Z., Pan Z. Geopolymer concrete at ambient and elevated temperatures: Recent developments
and challenges. NED University Journal of Research (Special issue of keynote papers presented
on the First South Asia Conference on Earthquake Engineering), 2019, 16(3): 113-128.

GEOPOLYMER CONCRETE AT AMBIENT AND ELEVATED


TEMPERATURES: RECENT DEVELOPMENTS AND CHALLENGES
Zhong Tao, Zhu Pan
Centre for Infrastructure Engineering, Western Sydney University, Penrith, NSW 2751, Australia
z.tao@westernsydney.edu.au, z.pan@westernsydney.edu.au

ABSTRACT
Geopolymer concrete (GPC) is attracting increasing attention due to its potential to replace ordinary
Portland cement (OPC) concrete to reduce the carbon dioxide emissions from cement production. Although
extensive research has been conducted in this area since the 1970s, its applications in engineering practice
are still very limited. This paper briefly reviews the recent developments of GPC and its properties at
ambient and elevated temperatures. It is found that GPC generally exhibits comparative properties to OPC
concrete at ambient temperature. In contrast, GPC often has better fire performance and superior durability.
Therefore, GPC could be advantageously used to improve the fire performance and durability of buildings
and infrastructure. While a lot of aluminosilicate raw materials (i.e., fly ash) are low cost, the prohibitive
costs of laboratory grade activators (i.e., sodium hydroxide and sodium silicate) greatly limit the widespread
application of GPC. Therefore, there is a need to develop low-cost geopolymer concrete. Meanwhile, a few
other challenges should also be overcome, such as difficulties to achieve consistent properties and to control
efflorescence of GPC. Further research is required at both material and structural levels to address these
issues. Particularly, there is a need to develop relevant building codes to promote the use of GPC in practice.

KEYWORDS
Geopolymer concrete, mechanical properties, fire performance, durability, challenges.

1. INTRODUCTION

Human activities over the past century have contributed significantly to the generation of greenhouse gases
(GHG), where about 70% GHG emissions are carbon dioxide (CO2) emissions. Global warming resulting
from the increased GHG concentration in the atmosphere could have many catastrophic consequences, such
as rising sea levels, extreme weather conditions, changing ecosystems and reduced food security (McCarthy
et al., 2010). Depending on the actual rate of GHG emissions, the global surface temperature on average
could rise a further 0.3 to 4.8 °C during the 21st century (IPCC, 2013). To curb the dangerous climate
change, the United Nations Framework Convention on Climate Change has set up targets of deep cuts in
GHG emissions to keep the temperature rise below 2 °C in this century.

Concrete is the most widely used construction material in the world, in which ordinary Portland cement
(OPC) is used as a binder material to bond aggregates together. Due to the increasing construction activities,
it is expected that global cement production will increase from 3.27 billion tonnes in 2010 to 4.83 billion
tonnes in 2030 (Statista, 2018). It is estimated that 0.87 tonne of CO2, including 0.53 tonne from limestone
decomposition and 0.34 tonne from energy consumption, are emitted from the production of every tonne
of cement (Provis and van Deventer, 2014). Because of the enormous volume, cement production accounts
for about 8% of global CO2 emissions (Andrew, 2018), and there is a need to take strategic actions to reduce
carbon emissions from this industry.

Great efforts have been made in the last few decades to develop alternative green binders to replace OPC
for reducing the environmental impact from construction. Geopolymer is considered as one of green
binders, which was first named by Davidovits in the late 1970s (Provis and van Deventer, 2014).
Geopolymer is produced by alkali activation of suitable aluminosilicate raw materials containing silicon
(Si) and aluminium (Al), such as metakaolin, fly ash (FA) and volcanic ash. The reaction produces
aluminate and silicate tetrahedra, which can combine to form tetrahedral frameworks linked by shared
oxygens (Davidovits, 2008).

It has been proved that geopolymer can be successfully used as a binder in GPC to produce structural
members (Tao et al., 2018). Although GPC was regarded as the “third generation cement” (after lime and
OPC) by some researchers (Li et al., 2010), its use in construction is still very limited. So far, it has been
mainly adopted in non-structural applications, such as pavements, sewer pipes, railway sleepers and
retailing walls (Gourley and Johnson, 2005; Aldred and Day, 2012). For example, GPC has been
successfully used to make pavements of the Westgate Freeway in Port Melbourne, Australia (van Deventer
et al., 2012). Another example is the use of 40,000 m3 GPC in making aircraft pavements and other facilities
at Brisbane West Wellcamp Airport, Australia (Glasby et al., 2015). So far, the Wellcamp Airport project
is still considered as the largest commercial application of GPC in the world. A structural use of Grade 40
MPa GPC was reported by Aldred and Day (2012), where a total of 33 floor beams were cast and used in
the Global Change Institute building at the University of Queensland, Australia.

Nowadays, geopolymer is still a very hot research topic, and publications in this field have increased
exponentially in the last two decades. Despite the extensive research, there are still a few major obstacles
for widespread application of this material. This paper presents an overview of recent developments of
geopolymer concrete and its limitations. The major challenges inherent in its utilisation and acceptance are
also discussed.

2. AN OVERVIEW OF GEOPOLYMER CONCRETE

2.1 Aluminosilicate Raw Materials

Geopolymer belongs to a subclass of alkali-activated materials (AAMs). According to Pacheco et al.


(2014), AAMs can be classified into three groups based on the microstructure of the binders: (1) high-
calcium AAM; (2) low-calcium AAM or geopolymer; and (3) hybrid alkaline cement. Ground granulated
blast-furnace slag (GGBS) is the most common prime material used to make high-calcium AAM. The main
reaction product of this type of AAM is a calcium silicate hydrate (C-A-S-H) gel, which is similar to the
gel obtained from the hydration of OPC, but with a lower Ca/Si ratio. To make geopolymer, however, low-
calcium Class F FA (CaO content is less than 10%) and metakaolin are the most popular prime materials.
FA is collected from the flue gas generated during the burning of coal in power plants. In commercial
production, FA is the preferred material than metakaolin, as FA is much cheaper than metakaolin (Pacheco
et al., 2014). The latter requires heating and milling to obtain a reactive powder. Furthermore, metakaolin
with plate-like particles requires a high amount of liquid to obtain a workable mixture, leading to low
mechanical properties. In geopolymer, the main reaction product is a three-dimensional alkaline
aluminosilicate hydrate (N-A-S-H) gel. It should be noted that the reaction of FA or metakaolin is relatively
slow at ambient temperature. Therefore, heat-curing at 60−120 °C is often required for geopolymer to
achieve a relatively high strength. If a small amount of calcium-rich material, i.e., slag, is added to the
mixture, the binder could obtain a high strength without additional heat curing. Blended with a small
amount of calcium-rich material, this type of AAM can be regarded as a hybrid cement, as the reaction
products are very complex, comprising a mix of C-A-S-H and N-A-S-H gels. It is worth mentioning that
FA rich in calcium (Class C) can also be activated by alkali to obtain high-calcium AAM (Chindaprasirt et
al., 2007). Since this paper mainly focuses on geopolymer, the default type of FA is Class F, unless
otherwise specified.

High-calcium AAM based on slag was invented as early as in the 1940s, which was much earlier than
geopolymer (Pacheco et al., 2014). Since slag is more reactive than FA or metakaolin, alkali-activated slag
normally can obtain much higher strength than geopolymer (Li et al., 2010). However, alkali-activated slag
usually sets very quickly, accompanied by high drying shrinkage. Meanwhile, the long-term performance
of alkali-activated slag could deteriorate, whereas the performance of geopolymer is more stable
(Wardhono et al., 2017). For these reasons, there is an increasing interest in investigating FA-based
geopolymer. It should be noted that geopolymer was originally used to describe low-calcium or calcium-
free alkali-activated systems only. But this term is gaining growing acceptance in the last few decades.
Despite the ongoing debate, the term “geopolymer” is now applied by many researchers to hybrid cement
or even high-calcium AAM. Unless otherwise specified, both low-calcium AAM and hybrid alkaline
cement are referred to as geopolymers in this paper.

2.2 Activators

The formation of geopolymer comprises three main steps (Davidovits, 2008): (1) dissolution of Si and Al
from the source material through the action of activator; (2) formation of oligomers consisting of polymeric
bonds of Si–O–Si and/or Si–O–Al type; and (3) setting or polycondensation of the oligomers to form an
inorganic polymeric structure. The presence of an alkali metal salt and/or hydroxide is essential for
dissolution of silica and alumina in the prime material under highly alkaline conditions, and for the catalysis
of the condensation reaction. The most common activators used are sodium silicate (Na2SiO3) and sodium
hydroxide (NaOH). Although potassium silicate and potassium hydroxide can also be used, it is more cost-
effective to use Na2SiO3 and NaOH to develop geopolymer. It is found that the combined use of Na2SiO3
and NaOH solutions can allow geopolymer to develop a higher mechanical strength compared with the use
of Na2SiO3 or NaOH alone (Phoo-ngernkham et al., 2017). For example, Fernandez-Jimenez and Palomo
(2003) reported a strength of 40 MPa for metakaolin-based geoplymer activated by NaOH after one day of
heat curing. When Na2SiO3 was added, the corresponding strength increased to 90 MPa. The NaOH solution
(normally 8−12 M) assists the dissolution of Si and Al from the precursor, whereas the additional silicate
species in the system can promote the geopolymerisation reaction as most aluminosilicate raw materials
cannot supply enough Si to the solution to start the geopolymerisation (Xu and van Deventer, 2000).

If the alkalinity of the NaOH solution is too high, the polymerisation process will be hindered, leading to
poor mechanical strength (Komnitsas and Zaharaki, 2007). Likewise, excessive Na2SiO3 inhibits the
geopolymerisation due to the formation of aluminosilicate gel precipitation, which obstructs the leaching
out of Si and Al from the source material (Leong et al., 2016). Although dosage of the activator and ratio
of Na2SiO3/NaOH significantly affect the cost, workability and strength development of GPC, the optimal
values vary with the nature of FA, type and amount of additives, and curing conditions (Pacheco et al.,
2014). Often trial tests are required to determine the optimal values.

2.3 Mixing Procedure

The mixing procedure used for GPC is very similar to that of normal concrete. Normally, dry materials,
including binders (i.e., FA and/or slag), sand, and coarse aggregate, are mixed together first, followed by
adding activator solution into the mixer (Cao et al., 2018). To achieve a desirable workability, extra water
is often added to the mixture to adjust the slump of fresh GPC (Hardjito et al., 2004). When NaOH pellets
are dissolved in water, a large amount of heat will be generated. Therefore, the NaOH solution is normally
prepared 24 h in advance to allow it to cool down. This will avoid the possible uncontrolled acceleration of
setting of the geopolymer (Phoo-ngernkham et al., 2017).

2.4 Curing Conditions

Due to the low reactivity of FA and metakaolin, (electrical or steam) heat curing is required for low-calcium
geopolymer to set and to obtain a high early strength. The geopolymerisation process can be greatly
promoted if the binder is cured at a temperature between 60 and 90 °C for 8−24 h (Provis and van Deventer,
2014). However, curing at a too high temperature or for a much longer period of time might weaken the
microstructure and/or lead to excessive shrinkage and cracking, causing a decrease in strength (van
Jaarsveld et al., 2002).

Unlike OPC, water does not directly participate in chemical reaction of geopolymer (Davidovits, 2008).
However, water serves as a medium to transport dissolved substances to allow the progress of
geopolymerisation (Zuhua et al., 2009). Meanwhile, adding an adequate amount of water is essential for
obtaining workable geopolymer. Because of the role of water in geopolymer, it is not necessary to cure
GPC in water. Instead, water curing leads to alkali leaching and lower compressive strength (Pacheco et
al., 2014). However, sealed curing to avoid moisture loss will be very helpful to obtain a dense and durable
matrix (Provis and van Deventer, 2014).

Microwave curing has also been explored by some researchers to cure geopolymer paste or mortar
(Chindaprasirt et al., 2013; Graytee et al., 2018). For conventional oven-curing, heat is transferred from the
exterior to the interior and this process takes a long period. In contrast, microwave curing can lead to
uniform and fast heating through interaction between polar molecules and microwave electric fields. The
microwave curing can be applied solely for a short period, i.e., ≤ 60 min (Graytee et al., 2018).
Alternatively, it can be applied in combination with the conventional heat curing to reduce the curing time
(Chindaprasirt et al., 2013). It should be noted that this technique has not been applied to GPC, and the
applicability to large-size structural members is yet to be proved.

2.5 Environmental Benefits

The research on geopolymer and GPC has been largely driven by the environmental benefits of geopolymer
binders (Pacheco et al., 2014). In general, researchers conclude that the use of GPC to replace OPC concrete
can reduce the carbon emissions. However, there is no consensus on the amount of savings, which greatly
depend on the mix design, particularly the amount of activators used. It is reported that Na2SiO3 has a high
carbon footprint, followed by NaOH. Habert et al. (2011) conducted a life cycle assessment for a typical
FA-based GPC mix with 17 kg/m3 NaOH powder and 103 kg/m3 Na2SiO3 solution. It is estimated that the
equivalent CO2 emissions are 306 kg for producing a cubic metre OPC concrete, whereas the corresponding
value is 169 kg for GPC, representing a saving of 45%. However, Turner and Collins (2013) reported that
the carbon footprint of GPC is only about 9% less than that of OPC. Turner and Collins (2013) assumed a
GPC mix (heat-cured at 60 °C for 24 h) containing 41 kg/m3 NaOH powder and 103 kg/m3 Na2SiO3
solution. Since ongoing efforts are being made to minimise the consumption of activators, further research
is required to justify the environmental benefits of those new types of GPC.

3. RECENT DEVELOPMENTS OF GEOPOLYMER CONCRETE

3.1 Geopolymer Concrete Cured at Ambient Temperature

Due to the low reactivity of Class F FA, heat curing is necessary for FA-based geopolymers to develop
reasonable strength. However, difficulties arise for in-situ applications to use this curing method.
Meanwhile, heat curing increases energy consumption and attracts a significant amount of investment for
space and facility (Pacheco et al., 2014). In the last two decades, extensive research has been conducted to
develop GPC to be cured at ambient temperature (around 20−25 °C). This can be achieved by blending fly
ash with high-calcium materials, such as slag (Deb et al., 2014), OPC (Nath and Sarker, 2015) and calcium
aluminate cement (Cao et al., 2018). The addition of calcium-rich material affects structural reorganisation
of the binder by increasing the formation of C-S-H gel, which in turn accelerates the strength development,
especially in the early age. Although it is debatable if the hybrid cements can still be called “geopolymer”,
they are more readily accepted by the construction industry (Habert and Ouellet-Plamondon, 2016) and
heralded as the “binders of the future” (Pacheco et al., 2014).

Previous studies demonstrate that the behaviour of hybrid cements are similar to that of OPC in terms of
workability, setting time and strength development if adequate amount of additives are included (Shang et
al., 2018). However, too much addition might lead to fast setting, poor workability and high shrinkage. The
suggested replacement ratios of FA are normally 20-30% for slag (Shang et al., 2018), 5−20% for OPC
(Nath and Sarker, 2015; Mehta and Siddique, 2017), and 5−10% for calcium aluminate cement (Cao et al.,
2018), respectively.

3.2 One-part Geopolymer Concrete

Conventional geopolymers are made based on a two-part system, i.e., solid materials and liquid alkali
solution (Davidovits, 2008). The alkali solutions are viscous and corrosive, which are difficult to handle
and not user-friendly (Luukkonen et al., 2017). Inspired by the production of OPC concrete, many
researchers have focused on the development of one-part “just add water” GPC in the last decade. Different
from the two-part system, one-part GPC is prepared by mixing water with a dry mix, consisting
aluminosilicate raw material, solid activator and other admixtures (Luukkonen et al., 2017).

In early efforts, kaolinite and red mud were often calcined with powdered hydroxide at temperatures
ranging from 550 to 1150 °C to make one-part mix binders (Feng et al., 2012). However, the high
temperature calcination greatly offsets the environmental benefits of one-part GPC and limits its
commercial applications. More recently, researchers developed FA-based one-part geopolymers by mixing
dry FA with various solid activators, such as NaOH and Na2SiO3. To avoid heat-curing, slag, hydrated lime
and OPC might also be added into the mixture (Nematollahi et al., 2015).

However, most reported one-part GPCs have little commercial value because of the use of a large amount
of activators. Indeed, there are a few promising trial mixes. For example, Nematollahi et al. (2015)
developed a mix containing 50 wt% fly ash and 50 wt% slag, and the activator used was anhydrous sodium
metasilicate powder and the dosage was 8 wt% of the total binder. The water/binder (w/b) ratio of the mix
was chosen as 0.35. This geopolymer mix achieved a compressive strength of 35.2 MPa at 28 days.
Meanwhile, reasonable rheology was also achieved. More recently, Askarian et al. (2018) developed a new
type of one-part GPC with a w/b ratio of 0.3, where solid potassium carbonate at a dosage of 7.5 wt% of
the total precursor was used as the activator. The geopolymer raw materials contained 10 wt% slag and 90
wt% FA. Meanwhile, the inclusion of OPC in the binder ranged from 10% to 60%. As the OPC replacement
ratio increased from 10% to 60%, the 28-day compressive strength increased from 33.4 to 55.0 MPa.
However, the slump decreased from 105 mm to 30 mm accordingly. It seems that the addition of 10% to
30% OPC is optimal to achieve a balance between strength and workability. In general, current studies on
one-part GPC have mainly focused on the development of mixtures to achieve reasonable strength. More
research is still required to develop commercially viable mix design. Meanwhile, attentions should be paid
to the material durability and structural performance of concrete members as well.
3.3 Geopolymer Concrete with High Flowability

Self-consolidating concrete (SCC) based on OPC has been widely used in practice to eliminate the need for
vibration. Modern superplasticiser plays a key role in achieving high flowability for SCC. Fresh
geopolymer, however, normally is viscous due to the use of alkaline solution (Xie and Kayali, 2016).
Unfortunately, current commercial superplasticisers are generally not very effective to improve the
workability of geopolymer since the molecules of the superplasticisers become instable in an environment
of high alkalinity (Xie and Kayali, 2016). Polycarboxylate-based superplasticiser worked better for Class
C FA-based GPC and naphthalene-based superplasticiser was more suitable for Class F FA-based GPC.

Research efforts to develop GPC with high flowability are still very scarce. Apart from the use of
superplasticisers, workability of GPC is normally controlled by adjusting the water content and the amount
and concentration of the alkaline solution (Hardjito et al., 2004). Nagaraj and Babu (2018) developed self-
compacting alkaline activated concrete using NaOH and Na2SiO3 solutions, where the precursor contained
75 wt% slag and 25 wt% fly ash. In their mix design, no superplasticisers were used due to their
ineffectiveness in trial tests. Instead, a very high amount of extra water of around 150 kg/m3 was added to
achieve a good mixture workability. Meanwhile, the amount of Na2SiO3 solution used in the mixes was
also very high, which is close to 200 kg/m3. Due to the large proportion of slag, the obtained compressive
strengths at 28 days are still reasonable, ranging from 39.5 to 59.0 MPa when the NaOH concentration is
in the range of 8−12 M. Obviously, a breakthrough in the development of superplasticiser is required for
making cost-effective self-consolidating GPC with a large volume of FA.

3.4 Low-cost Geopolymer Concrete

The environmental benefits of GPC for replacing OPC have been widely recognised, but its economic
benefits are less justifiable. While fly ash itself is low cost, the costs of laboratory grade NaOH and Na2SiO3
are very high. This greatly limits the widespread application of GPC. McLellan et al. (2011) did a cost
analysis for four typical mix design of GPC. They found that only one GPC mix can save a cost of 7%
compared with the OPC counterpart (Grade C40). The NaOH content in that particular GPC mix is 7.9 wt%
of fly ash and the corresponding ratio is 2 wt% for the sodium silicate solid. In contrast, the costs of the
three other GPC mixes are 11−39% higher than the cost of the OPC counterpart because of the increased
amount of activators or the addition of a large quantity of silica fume in the mixture. Obviously, it would
be more cost-effective if less alkali activators were used. As reported by Glasby et al. (2015), the GPC
developed for the Wellcamp Airport project used 415 kg/m3 aluminosilicate precursor (slag+FA) and 37
kg/m3 activator (solid content). Although details of the proprietary GPC mix were not given, the reported
activator to binder ratio of 8.9% can be considered as a benchmark for developing future low-cost GPC.

Since water glass (Na2SiO3 solution) contributes significantly to the cost of GPC, some researchers recently
tried to use home-made water glass to replace commercial water glass. In the laboratory, water glass was
made by dissolving silica-rich materials, such as glass powder (Torres-Carrasco and Puertas, 2014), rice
husk ash (Tchakoute et al., 2016), and silica fume (Assi et al., 2016), in 10−14 M NaOH solution at elevated
temperatures ranging from 75 to 100 °C. Constant magnetic stirring was employed through this process for
2−12 h. Due to the high amount of impurities in rice husk ash and glass powder, the home-made water glass
was found to be less effective than commercial water glass (Torres-Carrasco and Puertas, 2014), but the
glass powder-based water glass performed better than the rice husk ash-based water glass. On the other
hand, Assi et al. (2016) reported that silica fume-based water glass could perform even better than
commercial water glass. It could be attributed to the high purity of silica fume and the small particle sizes.
These studies highlight a very important research direction to develop home-made water glass to produce
cost-effective GPC.
4. BEHAVIOUR OF GEOPOLYMER CONCRETE AT AMBIENT TEMPERATURE

4.1 Fresh properties

The workability of GPC is affected by many factors, such as the shape and particle size distribution of the
prime materials, concentration of NaOH solution, the ratio of Na2SiO3/NaOH, type and amount of
superplasticiser, etc (Pacheco et al., 2014). Therefore, trial tests are often necessary for obtaining a mix
with optimal workability. Although extra water can be used to adjust the workability to obtain a slump
value up to 250 mm, it also leads to significant strength reduction. Fang et al. (2018) suggested a criterion
to define the workability of AAM systems: highly workable concrete with a slump of 90 mm or above,
medium workable concrete with a slump between 50 and 90 mm; and low workable concrete with a slump
lower than 50 mm. Due to the ineffectiveness of superplasticisers on GPC, the workability of GPC is
normally inferior to that of OPC concrete. For the same reason, GPC normally is not susceptible to
segregation and has higher cohesiveness than the OPC concrete (Pacheco et al., 2014).

Setting time of geopolymer normally increases as the SiO2/Al2O3 molar ratio (2.5-5.0) increases in the
initial mixture (Komnitsas and Zaharaki, 2007). Meanwhile, the setting time decreases as the pH of the
activating solution increases (Khale and Chaudhary, 2007). Furthermore, an increased temperature greatly
shortens the setting time, as Si and Al are easier to dissolve from the source material (Pacheco et al., 2014).
For Class F FA-based GPC, the setting time could be longer than 1 day at temperatures lower than 30 °C.
This issue can be resolved by blending FA with other calcium-rich compounds, such as slag or OPC.
However, this may lead to flash setting. Assi et al. (2018) reported that the initial and final setting times
were improved by 100% and 160%, respectively, when 3% of sucrose (sugar) was added. Meanwhile, the
compressive strength was not obviously affected. It seems that sucrose could be used to retard the initial
setting of some ambient-cured GPC.

4.2 Mechanical properties

Many factors can affect the mechanical properties of GPC, including reactivity of the prime material,
activator type and concentration, mix design and curing conditions (Pacheco et al., 2014). For example, a
prime material with a very high Al2O3 content normally produces a low-strength GPC. In the open literature,
the reported compressive strength of GPC normally ranges from 30 to 80 MPa (Hardjito et al., 2004). For
geopolymer mortars, Atiş et al. (2015) reported a compressive strength of 120 MPa for a geopolymer
mixture activated with 14% NaOH and cured at 115 °C for 24 h. Ambily et al. (2014) reported a very high
strength of up to 175 MPa for ambient cured ultra-high-performance geopolymer concrete, very little (33%)
or no FA, however, was used in the binder. Instead, the slag content in the binder ranged from 43% to 76%
and the addition of silica fume was as high as 22%.

The concrete density mainly depends on the density of aggregates used in the mixture. For this reason, the
density of normal GPC (2300−2400 kg/m3) is comparable to that of OPC concrete (Farhan et al. 2019).
Ding et al. (2016) reported that the elastic modulus of N-A-S-H gels in geopolymer is lower than that of C-
S-H gel. This might explain the widely reported finding that the elastic modulus (Ec) of heat-cured GPC is
lower than that of OPC. Compared with OPC, the reduction in Ec for GPC ranges from 10% to 29% (Nath
and Sarker, 2017). By testing ambient-cured GPCs blended with slag, OPC or Ca(OH)2, Nath and Sarker
(2017) found that ambient-cured GPC had a similar Ec compared with heat-cured GPC, and existing design
codes developed for OPC concrete normally give unconservative predictions of Ec for GPC. However, Cao
et al. (2018) reported that the Australian standard AS 3600 can give reasonable predictions of Ec for GPC
blended with CAC. Further research is required to confirm this finding.
Similar to OPC concrete, GPC has relatively low splitting tensile strength (fct) and flexural strength ( , ).
In general, fct and , of GPC increase with increasing compressive strength (Pacheco et al., 2014).
Heat-cured GPC generally has higher fct than the OPC concrete with a similar compressive strength, but
ambient-cured GPC and the OPC counterpart have similar fct–values (Deb et al., 2014). Regardless of the
curing condition, GPC normally has higher , than the OPC concrete with a similar (Deb et al., 2014;
Nath and Sarker, 2017). To improve the tensile and flexural performance of GPC, various organic or
inorganic fibres (i.e., PVA, PP, steel, E-glass and carbon) can be added into the mixture (Pacheco et al.,
2014).

4.3 Thermal properties

FA-based geopolymer paste generally has lower thermal conductivity and specific heat than the OPC paste.
Snell et al. (2017) reported thermal conductivities (λc) of 1.07 and 0.57 W/mK at room temperature for
OPC and geopolymer pastes, respectively. When the paste content in concrete is kept at 25%, λc-values of
1.56 and 1.26 W/mK were found for OPC and geopolymer concretes, respectively. The reported values of
specific heat (cp) are 877.9, 730.1, 718.4, and 681.4 J/kg⋅K for the OPC paste, geopolymer paste, OPC
concrete and GPC, respectively. The lower thermal conductivity of GPC can be beneficial when GPC is
used in construction for energy conservation. The coefficient of thermal expansion (CTE) for OPC concrete
ranges from 8 to 12 µε/K. A high value of 16.6 µε/K was reported by Ma and Dehn (2017) for heat-cured
GPC concrete. In contrast, Junaid (2017) reported values ranging from 10.3 to 10.9 µε/K, which are similar
to that of OPC concrete. Since thermal effects are important for structures, further research is required to
clarify the inconsistency in CTE of GPC concrete.

4.4 Drying shrinkage and creep

It is generally accepted that heat-cured GPC has negligible drying shrinkage because most of water has
evaporated during the heating procedure (Pacheco et al., 2014). The measured maximum shrinkage of heat
cured GPC was only about 100 µε within one year, as reported by Hardjito et al. (2004). Instead, ambient-
cured GPC has much larger drying shrinkage than heat-cured GPC since water in the former takes a long
time to evaporate. The other influencing factor is the replacement ratio of FA. Shang et al. (2018) found
that drying shrinkage of geopolymer mortar increases with increasing slag content in the mixture. This
could be due to the change in microstructure and pore size distribution. When the slag percentage is less
than 20%, the geopolymer mortar had a drying shrinkage similar to the OPC counterpart. If excessive slag
was added, however, the mix experienced a relatively high shrinkage.

Creep of GPC has seldom been studied in the past. According to Pacheco et al. (2014), the final specific
creep of heat-cured GPC (fc′= 40 − 67 MPa) ranges from 15 to 29 µε/MPa after one-year loading. But for
60 MPa OPC concrete, the corresponding value is about 50 to 60 µε/MPa. It seems the creep strain of heat-
cured GPC is smaller than that of the OPC concrete. This might be explained by the theory of “micro-
aggregates” since the FA particles are not fully dissolved in the binder system (Pacheco et al., 2014). No
test data is available for the creep behaviour of ambient-cured GPC. Although Un et al. (2015) reported
long-term deflection of two beams made with alkali-activated cement concrete, the binder consisted of 380
kg/m3 slag and only 20 kg/m3 FA. Therefore, the concrete should be considered as alkali-activated slag
concrete rather than GPC. With this mix design, the two beams experienced large drying shrinkage and
long-term deflection.

4.5 Bond Strength between GPC and Reinforcement

The bond strength of GPC with steel reinforcement is generally higher than that of the OPC concrete. For
an equivalent compressive strength, Castel and Foster (2015) reported an average increase in bond strength
of 10% for heat-cured GPC consisting of 85.2% FA and 14.8% slag in the binder. The reported increase in
bond strength is even higher in tests conducted by Sarker (2011) for heat-cured FA-based GPC. Similar to
the OPC concrete, bond strength of GPC increases with increasing concrete cover and compressive strength
(Sarker, 2011). Meanwhile, deformed bars in GPC can develop much higher bond strength than smooth
bars (Castel and Foster, 2015).

4.6 Static Behaviour of GPC Members

Compared with the extensive research conducted on geopolymer and GPC materials, much less studies
have been conducted on GPC members, such as beams (Sumajouw and Rangan, 2006; Yost et al. 2013)
and columns (Sumajouw et al., 2007). In general, behaviour of reinforced GPC members is found to be
very similar to that of conventional reinforced OPC concrete members. Depending on the properties of the
GPC, some minor differences in behaviour were sometimes also reported. For example, Yost et al. (2013)
reported a more brittle crushing failure of GPC than OPC concrete in beams. Meanwhile, higher first crack
load, mid-span deflection and ultimate load, as well as smaller crack width, were also reported by other
researchers for GPC beams. More details can be found in the comprehensive literature review conducted
by Mo et al. (2016). In general, existing design codes, such as AS3600 and ACI 318, can be conservatively
used to predict the ultimate strength of reinforced GPC members (Mo et al., 2016), although further research
is still required to develop more economic and effective design methods for those members.

Ozbakkaloglu and Xie (2016) conducted compressive tests on 36 concrete-filled square fibre-reinforced
polymer tubes. They found that the use of GPC or OPC concrete had little influence on the compressive
strength of the composite columns. However, the GPC columns developed a lower ultimate axial strain
than the reference OPC concrete samples. Meanwhile, the OPC concrete samples had a marked plateau in
the transition regions of their stress−strain curves, which was not found in the GPC columns. This behaviour
difference was explained by the fact that the OPC concrete developed a much larger shrinkage than the
heat-cured GPC.

More recently, Tao et al. (2018) reported 5 tests on circular concrete-filled steel tubular (CFST) stub
columns. The test results demonstrate that GPC-filled steel tubular (GCFST) columns have similar ambient
temperature behaviour compared with the conventional CFST counterparts, which can be seen from the
comparison shown in Fig. 1. Due to the difference in concrete strength, the ultimate strengths of the three
composite columns are slightly different. However, the shapes of the axial load versus axial strain (N−ε)
curves for the two GPC specimens (CFT-GPC2 and CFT-GPC3) are similar to that of the conventional
CFST column (CFT-OPC). Meanwhile, the ultimate axial strains corresponding to the ultimate strengths
for specimens CFT-GPC2 and CFT-GPC3 are about 25% smaller than the corresponding strain (5860 µε)
of CFT-OPC. This is consistent with the observations reported by Ozbakkaloglu and Xie (2016). In general,
finite element models and design equations developed for conventional CFST columns can also be used to
predict N−ε curves and strength of GCFST columns, respectively, and reasonable prediction accuracy has
been achieved.

4.7 Cyclic Performance of GPC Members

Very limited studies have been conducted to study the cyclic performance of GPC members. Philip et al.
(2015) found that GPC could be successfully used to retrofit seismically damaged concrete frames. The
effectiveness was verified by shake table tests. Raj et al. (2016) reported test results of six GPC beam-
column joints and four conventional OPC concrete joints tested under reverse cyclic loading. The FA-based
GPC was heat-cured and achieved a compressive strength of 34 MPa. It was found that the two types of
joints had very similar cyclic behaviour, first crack loads and ultimate loads. However, the energy
absorption capacity of GPC joints was 39% higher than that of the OPC concrete joints. Meanwhile, the
GPC joints had slower deterioration in stiffness and increased displacement ductility.

5. BEHAVIOUR OF GEOPOLYMER CONCRETE AT ELEVATED TEMPERATURES

5.1 Spalling of GPC

OPC concrete is susceptible to spalling (often in an explosive way) during fire exposure, which is the
breaking away of pieces or layers of concrete due to the build-up of pore pressure or thermal stress (Ali et
al., 2017). In general, GPC is less susceptible to spalling compared with OPC concrete. Zhao and Sanjayan
(2011) tested GPC (fc′ = 37−98 MPa) and OPC concrete (fc′ = 42−110 MPa) cylinders by heating one end
to about 850 °C in approximately 2 min. They found that all GPC cylinders did not spall regardless of
strength. In contrast, OPC concrete with a fc′ of 42 or 63 MPa showed minor spalling, and high-strength
OPC concrete (fc′ = 82 or 110 MPa) exhibited severe or even explosive spalling. GPC cylinders were also
tested by Sarker et al. (2014) and Gluth et al. (2016) by exposing to ISO 834 standard fire, and no spalling
was found. The superior performance of GPC was explained by its relatively high permeability (Zhao and
Sanjayan, 2011), low amount of chemically bound water (Gluth et al., 2016), and high thermal compatibility
between aggregate and geopolymer paste (Ali et al., 2017). It should be noted that Pan et al. (2012) reported
severe spalling of GPC cylinders when the maximum aggregate size was 10 mm or less.

More recently, Ali et al. (2017) reported heat-cured FA-based GPC wall panels (fc′ = 64 MPa) subjected to
a hydrocarbon fire for 2 h. When concrete cylinders did not experience spalling, the panels experienced
some minor surface spalling up to a maximum depth of only 2 mm, and the weight loss due to spalling was
less than 2%. In contrast, OPC concrete panels with a thickness of 200 mm lost up to 191 mm after exposure
to a hydrocarbon fire for 1 h (Ali et al., 2017). It should be pointed out that no spalling test results have
been reported on ambient-cured GPC blended with calcium-rich material(s). It would be interesting to know
if such ambient-cured GPC is susceptible to spalling or not.

5.2 Hot Strength of GPC

Although OPC concrete can be rated as a fire-resistant material, its strength and stiffness deteriorate
significantly during fire exposure. According to Eurocode 2, the hot strength of OPC concrete at 800 °C
reduces to 27% of its initial strength at room temperature, as shown in Fig. 2. The strength deterioration is
mainly due to the dehydration and decomposition of hydration products in the OPC paste, as well as thermal
incompatibility between aggregates and the paste (Shaikh and Vimonsatit, 2015). For heat-cured GPC, its
three-dimensional N-A-S-H type gel is very thermally stable (Pan et al., 2018b), leading to high strength
retention for GPC, as shown in Fig. 2. A strength decrease is normally observed at a temperature (T) up to
400 °C due to the thermal incompatibility between aggregates and geopolymer paste (Shaikh and
Vimonsatit, 2015). Beyond this temperature, further geopolymerisation could occur, leading to an increase
in compressive strength.

However, the strength retention of GPC is greatly affected by the concentration of NaOH. According to the
tests conducted by Shaikh and Vimonsatit (2015), the concrete strength increased from about 35 MPa to 70
MPa as the NaOH concentration increased from 10 M to 16 M. However, fc′of GPC activated by 10 M or
13 M NaOH reached about 53 MPa at 800 °C, which is higher than the corresponding value of 38 MPa for
the GPC activated by 16 M NaOH. It seems that the geopolymerisation potential during fire exposure
reduces for GPC with a higher degree of geopolymerisation before fire exposure. The initial strength of
GPC also affects the strength retention of GPC at elevated temperatures. Junaid et al. (2017) reported
stress−strain (σ−ε) curves of two GPC mixes at different temperatures up to 800 °C. The mix proportions
of the two mixes are virtually the same except that extra water of 32.3 kg/m3 was added to mix M7. Values
of fc′ for M7 and M14 (without extra water) are 40.2 and 60.7 MPa. At 800 °C, fc′ of M14 decreased to
about 27 MPa, which is lower than the corresponding value of 37 MPa obtained for M7. The reason for this
is still not clear.

1800 1.5
OPC concrete GPC (Shaikh and Vimonsatit, 2015)

Hot strength ratio %


1500 1.25 GPC (Cao et al., 2017)
OPC (Eurocode 2) 10 M NaOH
Axial load N (kN)

1200 1
GPC
900 0.75
εy = 1680 µε
13 M NaOH
600 CFT-OPC 0.5
CFT-GPC2
300 0.25
CFT-GPC3
0 0
0 10000 20000 30000 40000 0 200 400 600 800
Axial strain ε (×10−6) Temperature (oC)

Figure 1 Comparison between N−ε curves of Figure 2 Strength retention of GPC and OPC concrete
GCFST and conventional CFST columns at elevated temperatures

For ambient-cured GPC, it might still have better strength retention capacity compared with OPC. Cao et
al. (2017) developed GPC by incorporating into 10% CAC in the binder and the NaOH concentration is 14
M. The influence of T on this type of ambient-cured GPC is shown in Fig. 2, which has comparable strength
retention capacity as the heat-cured GPC (activated by 13 M NaOH) tested by Shaikh and Vimonsatit
(2015). More recently, Pan et al. (2018b) reported tested results of ambient-cured geopolymer mortars. A
strength loss of 50% was reported for OPC mortar at 600 °C. For geopolymer mortars with 10 and 50%
slag in the binders, the corresponding strength decreases are 18% and 37%, respectively. Due to the limited
test data, it is still not possible to develop reliable models to predict hot strength or σ−ε curves of GPC.
Further research is required in this area.

5.3 Thermal Properties of Geopolymer Binders

Pan et al. (2018a) reported thermal properties of various geopolymer binders, where FA/slag ratios are
100/0 (F100), 90/10 (F90S10) and 50/50 (F50S50) wt%, respectively. The FA-based geopolymer paste was
cured at 60 °C for 24 h, whereas other geopolymer pastes and OPC reference paste were ambient-cured.
The thermal conductivities of different binders are shown in Fig. 3 as a function of T. For the reference
OPC sample, the thermal conductivity decreases with increasing temperature up to 400 °C and then remains
stable. The decrease in thermal conductivity is due to moisture loss. For mixes F100 and F90S10, thermal
conductivities increase with increasing temperature due to densification of the pore structure resulting from
further geopolymerisation. In contrast, thermal conductivities of F50S50 are relatively stable in the
temperature range from 20 to 600 °C.

The values of specific heat of different binders are shown in Fig. 4 as a function of T. In general, specific
heat increases with increasing T due to the moisture evaporation and/or chemical decomposition (Pan et
al., 2018a). The sharp decrease of specific heat for F100 in the temperature range from 200 to 300 °C is
due to the exothermic geopolymerisation.

5.4 Fire Performance of GPC Members

Sarker and Mcbeath (2015) and Luna-Galiano et al. (2015) conducted fire resistance tests on FA-based
GPC panels subjected to ISO 834 standard fire. Although the GPC panels showed excellent integrity and
higher residual load-carrying capacity, they might have higher temperature in the unheated face compared
with the OPC panel with a same thickness. This might be explained by the measured thermal conductivities
of geopolymer binders shown in Fig. 3. Although GPC might have a lower thermal conductivity at ambient
temperature, its thermal conductivity could soon exceed that of the OPC concrete at elevated temperatures.

1.2 3.5
OPC Fly ash OPC

Specific heat (MJ/m3K)


Thermal conductivity (W/mK)

90% fly ash/10% slag 2.8


0.9
Fly ash
2.1
0.6 90% fly ash/10% slag
1.4 50% fly ash/50% slag

0.3 50% fly ash/50% slag 0.7

0 0
0 200 400 600 0 200 400 600
Temperature (oC) Temperature (oC)
Figure 3 Thermal conductivity at high temperatures Figure 4 Specific heat at high temperatures

As mentioned in Section 5.2, GPC generally has a high strength retention capacity than OPC concrete.
Therefore, it is expected that the use of GPC could improve the fire performance of structural members.
Espinos et al. (2015) conducted a numerical analysis to study the behaviour of concrete-filled double-tube
columns in fire. The predicted fire resistance time for a typical OPC-filled double-tube column is 87 min.
In contrast, the predicted fire resistance time increases to 139 min when GPC is used to fill the ring between
the inner and the outer steel tubes. The increase in fire resistance is due to the delay in the temperature rise
of the inner tube for the use of GPC. Further tests are required to verify the simulation results.

In Australia, internal reinforcement is often used to improve the fire resistance of CFST columns. Tao et
al. (2018) proposed the concept to use GPC to replace conventional OPC concrete in CFST columns. Due
to the excellent fire resistance of GPC, the internal reinforcement in a CFST column might be eliminated.
To prove the concept, Tao et al. (2018) conducted fire tests on 5 CFST stub columns. The initial applied
load was 50% of the ultimate load measured at ambient temperature. Then the furnace temperature was
increased to 800 °C and kept at this temperature until the column failed. Fig. 5 compares the axial
deformation versus time curves for different specimens. The conventional CFST column (HOT-OPC) has
the worst fire resistance (fire resistance time tR = 36.7 min). In contrast, the heat-cured GCFST column
(HOT-GPC4) has the best fire resistance (tR = 85.4 min). Fire resistance of other ambient-cured GCFST
columns with 5-10% CAC in the binder decreases with increasing GPC strength. The fc′-values are 37.4,
58.6 and 64.4 MPa for specimens HOT-GPC1, HOT-GPC2, and HOT-GPC3, respectively.

6. CHALLENGES OF USING GEOPOLYMER CONCRETE

6.1 High Cost to Produce Geopolymer Concrete

Since OPC concrete is dominantly used in practice and already has a proven track record, additional efforts
should be made to prove the benefits of using GPC to replace OPC concrete. As construction industry is
very conservative to adopt new technologies to replace existing ones, environmental benefits themselves
are unlikely to motivate construction companies to adopt GPC. Although FA itself is relatively cheap, the
costs of laboratory grade activators, i.e. NaOH and Na2SiO3, are relatively expensive. Majidi (2009)
estimated that GPC products are 10-15% more expansive than OPC products. Shang et al. (2018) even
reported that GPC (CNY 1277.3-1302.3 per tonne) was 3.6 times more expansive than OPC concrete (CNY
285.5 per tonne) in China.

0
Deformation (mm)
-6

-12 HOT-OPC
HOT-GPC1
-18 HOT-GPC2
HOT-GPC3
-24 HOT-GPC4
-30
0 15 30 45 60 75 90
Time (min)

Figure 5 Axial deformation−time curves of CFST columns subjected to elevated temperatures

Since GPC has excellent chemical resistance and fire resistance, it would be beneficial to use GPC to replace
OPC in aggressive environments or in situations with a demand of high fire resistance. In these cases, it
might be more cost-effective for using GPC after considering repair and maintenance costs in the life cycle
of a structure or the saving of additional steel reinforcement or fire coating material. Meanwhile, there is a
need to develop GPC using less alkaline activators. By adding calcium-rich material, it is possible to reduce
the concentration of the alkaline solutions (Pacheco et al., 2014). Furthermore, there is a need to develop/use
new cost-effective activators to replace commercially available sodium silicate. Some recent works have
been reviewed in Subsection 3.4.

6.2 Difficulties to Achieve Consistent Properties

Although metakaolin-based geopolymers are easy to obtain controlled properties, they are not suitable to
be used as primary binders in GPC, as explained in Subsection 2.1. In contrast, FA is an industry by-product,
including a significant amount of impurities (Provis and van Deventer, 2014). The quality of FA is
dependent on the coal type and combustion furnace (Pacheco et al., 2014). The Si and Al contents in FA
can also vary significantly with Si varying from 40 to 60% and Al from 20 to 30%. The particles in FA are
also inhomogeneous, comprising both glassy and crystalline phases at different proportions (Khale and
Chaudhary, 2007). The crystalline phases, such as mullite and quartz, however, exhibit low relativity.
Meanwhile, the particle size distributions of fly ashes from different sources could be quite different.
Therefore, it is a challenge to obtain consistent properties when using different FA source materials. Instead,
trial tests are often required to check the potential of a particular FA to be alkali activated, and to obtain the
optimal mix design (Khale and Chaudhary, 2007).

Although FA could potentially be mechanically, thermally or chemically treated to improve its reactivity
(Tchadjie and Ekolu, 2018), it is normally not cost-effective to adopt this option. To copy with the
inconsistency issue of FA-based GPC, two options may be adopted. The first option is to use FA from a
known source with a proven track record in making GPC. This is a viable solution if a prefabrication plant
can be built near the particular coal-fired power station. The second option is to develop alkali cements
using binary, ternary or even quaternary blends. In this case, FA can be blended with other ingredients with
more consistent performance, such as metakaolin, slag and OPC.
6.3 Difficulties to Control Efflorescence

GPC could subject to severe efflorescence due to the high amount of alkali and soluble silicate, which
cannot be fully consumed (Kani et al., 2012). In most cases, efflorescence is due to the formation of white
crystalline sodium carbonates resulting from the reaction between atmospheric carbon dioxide and excess
alkali. Efflorescence is not only an aesthetic problem to GPC, but can also cause significant microstructure
damage due to increased pore pressure (Zhang et al., 2018).

To control efflorescence, the amount of activators used in GPC could be potentially reduced, which can
also help to reduce the cost of GPC. Kani et al. (2012) suggested the addition of 8% CAC to reduce the
mobility of alkalis and adoption of heat curing. Another control measure is to avoid the use of GPC in a
moisture environment or in contact with water.

6.4 Lack of Building Codes

GPC is a big family consisting many different types of materials. However, most developed GPC do not
have commercial value. There is an urgent need for researchers to focus on developing commercially-viable
mix design through engaging with end-users. Research can then be conducted in different areas for
developing building codes to promote the widespread use of low-cost GPCs. Suitable testing methods
should also be developed to test efflorescence, alkali-silica reaction and other properties of GPC.

6.5 Lack of Effective Superplasticiser

Another major challenge of using GPC is the lack of effective superplasticiser. Although the current
commercial superplasticisers are not very effective in improving the workability of GPC (Xie and Kayali,
2016), they are often still added to GPC at a large dosage to achieve the best workability. Chouhan et al.
(2018) recently developed rice husk-based superplasticiser by dissolving rice husk in boiled 12.5 M NaOH
solution. By adding 1% of the developed superplasticiser, the slump of GPC increased from 40 to 200 mm.
Although the results are very promising, more research should be conducted to verify the effectiveness of
the superplasticiser in other GPC mixes.

7. CONCLUDING REMARKS

Geopolymer concrete (GPC) generally exhibits comparative properties to OPC concrete at ambient
temperature, but often has better fire performance and superior durability. Therefore, GPC could be
advantageously used to improve the fire performance and durability of buildings and infrastructure.
Although extensive research has been conducted on geopolymer since the 1970s, its applications in
engineering practice are still very limited. The major challenge is the high cost of activators (i.e., sodium
hydroxide and sodium silicate), which make GPC less cost-effective. There is an urgent need for researchers
to focus on developing commercially-viable mix design through engaging with end-users. Meanwhile, a
few other challenges should also be overcome, such as the lack of effective superplasticiser and difficulties
to achieve consistent properties and to control efflorescence of GPC. Further research is required at both
material and structural levels to address these issues. Particularly, there is a need to develop relevant
building codes to promote the use of GPC in practice.

ACKNOWLEDGEMENTS

The authors are grateful for the financial support from Australian Research Council Linkage Grant No.
LP160101484.
REFERENCES

Ali, A.M., Sanjayan, J., and Guerrieri, M. (2017). “Performance of geopolymer high strength concrete wall
panels and cylinders when exposed to a hydrocarbon fire”. Construction and Building Materials, Vol.
137, pp 195-207.
Ambily, P.S., Ravisankar, K., Umarani, C., Dattatreya, J.K., and Iyer, N.R. (2014). “Development of ultra-
high-performance geopolymer concrete”. Magazine of Concrete Research, Vol. 66, No. 2, pp 82-89.
Andrew, R.M. (2018). “Global CO2 emissions from cement production”. Earth System Science Data, Vol.
10, No. 1, pp 195.
Askarian, M., Tao, Z., Adam, G., and Samali, B. (2018). “Mechanical properties of ambient cured one-part
hybrid OPC-geopolymer concrete”. Construction and Building Materials, Vol. 186, pp 330-337.
Assi, L.N., Deaver, E.E., ElBatanouny, M.K., & Ziehl, P. (2016). “Investigation of early compressive
strength of fly ash-based geopolymer concrete”. Construction and Building Materials, Vol. 112, pp
807-815.
Assi, L.N., Deaver, E.E., and Ziehl, P. (2018). “Using sucrose for improvement of initial and final setting
times of silica fume-based activating solution of fly ash geopolymer concrete”. Construction and
Building Materials, Vol. 191, pp 47-55.
Atiş, C.D., Görür, E.B., Karahan, O., Bilim, C., Ilkentapar, S., and Luga, E. (2015). “Very high strength
(120 MPa) class F fly ash geopolymer mortar activated at different NaOH amount, heat curing
temperature and heat curing duration”. Construction and building materials, Vol. 96, pp 673-678.
Cao, Y.F., Tao, Z., Pan, Z., and Wuhrer, R. (2018). “Effect of calcium aluminate cement on geopolymer
concrete cured at ambient temperature”. Construction and Building Materials, Vol. 191, pp 242-252.
Castel, A., and Foster, S.J. (2015). “Bond strength between blended slag and Class F fly ash geopolymer
concrete with steel reinforcement”. Cement and Concrete Research, Vol. 72, pp 48-53.
Chindaprasirt, P., Chareerat, T., and Sirivivatnanon, V. (2007). “Workability and strength of coarse high
calcium fly ash geopolymer”. Cement and concrete composites, Vol. 29, No. 3, pp 224-229.
Chindaprasirt, P., Rattanasak, U., and Taebuanhuad, S. (2013). “Role of microwave radiation in curing the
fly ash geopolymer”. Advanced Powder Technology, Vol. 24, No. 3, pp 703-707.
Chouhan, R.K., Mudgal, M., Bisarya, A., and Srivastava, A.K. (2018). “Rice-husk-based superplasticizer
to increase performance of fly ash geopolymer concrete”. Emerging Materials Research, Vol. 77, No.
3, pp 169-177.
Davidovits, J. (2008). “Geopolymer chemistry and applications”. Geopolymer Institute.
Deb, P.S., Nath, P., and Sarker, P.K. (2014). “The effects of ground granulated blast-furnace slag blending
with fly ash and activator content on the workability and strength properties of geopolymer concrete
cured at ambient temperature”. Materials & Design, Vol. 62, pp 32-39.
Ding, Y., Dai, J. G., and Shi, C. J. (2016). “Mechanical properties of alkali-activated concrete: A state-of-
the-art review”. Construction and Building Materials, Vol. 127, pp 68-79.
Espinos, A., Romero, M.L., Hospitaler, A., Pascual, A.M., and Albero, V. (2015). “Advanced materials for
concrete-filled tubular columns and connections”. Structures, Vol. 4, pp 105-113.
Fang, G., Ho, W.K., Tu, W., and Zhang, M. (2018). “Workability and mechanical properties of alkali-
activated fly ash-slag concrete cured at ambient temperature”. Construction and Building
Materials, Vol. 172, pp 476-487.
Farhan, N.A., Sheikh, M.N., and Hadi, M.N. (2019). “Investigation of engineering properties of normal and
high strength fly ash based geopolymer and alkali-activated slag concrete compared to ordinary
Portland cement concrete”. Construction and Building Materials, Vol. 196, pp 26-42.
Feng, D., Provis, J.L., and van Deventer, J.S.J. (2012). “Thermal activation of albite for the synthesis of
one-part mix geopolymers”. Journal of the American Ceramic Society, Vol. 95, No. 2, pp 565–572.
Fernández-Jiménez, A., and Palomo, A. (2003). “Characterisation of fly ashes. Potential reactivity as
alkaline cements”. Fuel, Vol. 82, No. 18, pp 2259-2265.
Glasby, T., Day, J., Genrich, R. and Aldred, J. (2015). "EFC geopolymer concrete aircraft pavements at
Brisbane west Wellcamp Airport", Concrete 2015 Conference, Melbourne, Australia.
Gluth, G.J., Rickard, W.D., Werner, S., and Pirskawetz, S. (2016). “Acoustic emission and microstructural
changes in fly ash geopolymer concretes exposed to simulated fire”. Materials and Structures, Vol. 49,
No. 12, pp 5243-5254.
Gourley, J.T., and Johnson, G.B. (2005). “Developments in geopolymer precast concrete”, World Congress
Geopolymer, Saint-Quentin, France, pp 139-143.
Graytee, A., Sanjayan, J.G., and Nazari, A. (2018). “Development of a high strength fly ash-based
geopolymer in short time by using microwave curing”. Ceramics International, Vol. 44, No. 7, pp
8216-8222.
Habert, G., De Lacaillerie, J.D.E., and Roussel, N. (2011). “An environmental evaluation of geopolymer
based concrete production: reviewing current research trends”. Journal of Cleaner Production, Vol.
19, No. 11, pp 1229-1238.
Habert, G., and Ouellet-Plamondon, C. (2016). “Recent update on the environmental impact of
geopolymers”. RILEM Technical Letters, Vol. 1, pp 17-23.
Hardjito, D., Wallah, S.E., Sumajouw, D.M., and Rangan, B.V. (2004). “On the development of fly ash-
based geopolymer concrete”. Materials Journal, Vol. 101, No. 6, pp 467-472.
IPCC, (2013). “Climate change 2013: The physical science basis -technical summary". Intergovernmental
Panel on Climate Change, pp 89–90.
James, A., and John, D. (2012). "Is geopolymer concrete a suitable alternative to traditional concrete?",
37th Conference on Our World in Concrete & Structures, Singapore, pp 1-14.
Junaid, M.T. (2017). “Deformational behaviour of fly-ash based geopolymer concrete at temperatures of
up to 150° C”, MATEC Web of Conferences, Vol. 120, EDP Sciences, pp 02020.
Junaid, M.T., Kayali, O., and Khennane, A. (2017). “Response of alkali activated low calcium fly-ash based
geopolymer concrete under compressive load at elevated temperatures”. Materials and Structures, Vol.
50, No. 1, pp 50.
Kani, E.N., Allahverdi, A., and Provis, J.L. (2012). “Efflorescence control in geopolymer binders based on
natural pozzolan”. Cement and Concrete Composites, Vol. 34, No. 1, pp 25-33.
Khale, D., and Chaudhary, R. (2007). “Mechanism of geopolymerization and factors influencing its
development: a review”. Journal of Materials Science, Vol. 42, No. 3, pp 729-746.
Komnitsas, K., and Zaharaki, D. (2007). “Geopolymerisation: A review and prospects for the minerals
industry”. Minerals Engineering, Vol. 20, No. 14, pp 1261-1277.
Leong, H.Y., Ong, D.E.L., Sanjayan, J.G., and Nazari, A. (2016). “The effect of different Na2O and K2O
ratios of alkali activator on compressive strength of fly ash based-geopolymer”. Construction and
Building Materials, Vol. 106, pp 500-511.
Li, C., Sun, H., and Li, L. (2010). “A review: The comparison between alkali-activated slag (Si+ Ca) and
metakaolin (Si+ Al) cements”. Cement and Concrete Research, Vol. 40, No. 9, pp 1341-1349.
Luna-Galiano, Y., Cornejo, A., Leiva, C., Vilches, L.F., and Fernández-Pereira, C. (2015). “Properties of
fly ash and metakaolín based geopolymer panels under fire resistance tests”. Materiales de
Construcción, Vol. 65, No. 319, pp 059.
Luukkonen, T., Abdollahnejad, Z., Yliniemi, J., Kinnunen, P., and Illikainen, M. (2017). “One-part alkali-
activated materials: A review”. Cement and Concrete Research, Vol. 103, pp 21-34.
Ma, J., and Dehn, F. (2017). “Investigations on the coefficient of thermal expansion of a low‐calcium fly
ash‐based geopolymer concrete”. Structural Concrete, Vol. 18, No. 5, pp 781-791.
Majidi, B. (2009). “Geopolymer technology, from fundamentals to advanced applications: a
review”. Materials Technology, Vol. 24, No. 2, pp 79-87.
McCarthy, M.P., Best, M.J., and Betts, R.A. (2010). “Climate change in cities due to global warming and
urban effects”. Geophysical Research Letters, Vol. 37, No. 9, pp 1-5.
McLellan, B.C., Williams, R.P., Lay, J., Van Riessen, A., and Corder, G.D. (2011). “Costs and carbon
emissions for geopolymer pastes in comparison to ordinary Portland cement”. Journal of Cleaner
Production, Vol. 19, No. 9-10, pp 1080-1090.
Mehta, A., and Siddique, R. (2017). “Properties of low-calcium fly ash based geopolymer concrete
incorporating OPC as partial replacement of fly ash”. Construction and Building Materials, Vol. 150,
pp 792-807.
Mo, K.H., Alengaram, U.J., and Jumaat, M.Z. (2016). “Structural performance of reinforced geopolymer
concrete members: A review”. Construction and Building Materials, Vol. 120, pp. 251-264.
Nagaraj V.K, and Babu, D.V. (2018). “Assessing the performance of molarity and alkaline activator ratio
on engineering properties of self-compacting alkaline activated concrete at ambient
temperature”. Journal of Building Engineering, Vol. 20, pp 137-155.
Nath, P., and 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”. Cement and Concrete
Composites, Vol. 55, pp 205-214.
Nath, P., and Sarker, P.K. (2017). “Flexural strength and elastic modulus of ambient-cured blended low-
calcium fly ash geopolymer concrete”. Construction and Building Materials, Vol. 130, pp 22-31.
Nematollahi, B., Sanjayan, J., and Shaikh, F.U.A. (2015). “Synthesis of heat and ambient cured one-part
geopolymer mixes with different grades of sodium silicate”. Ceramics International, Vol. 41, No. 4,
pp 5696-5704.
Ozbakkaloglu, T., and Xie, T. (2016). “Geopolymer concrete-filled FRP tubes: Behavior of circular and
square columns under axial compression”, Composites Part B: Engineering, Vol. 96, pp 215-230.
Pacheco, T., Fernando, L.J. and Leonelli, C. (2014) Handbook of Alkali-activated cements, Mortars and
Concretes, Woodhead Publishing.
Pan, Z., Sanjayan, J.G., and Kong, D.L. (2012). “Effect of aggregate size on spalling of geopolymer and
Portland cement concretes subjected to elevated temperatures”. Construction and Building
Materials, Vol. 36, pp 365-372.
Pan, Z., Tao, Z., Cao, Y.F., and Wuhrer, R. (2018a). “Measurement and prediction of thermal properties of
alkali-activated fly ash/slag binders at elevated temperatures”. Materials and Structures, Vol. 51, No.
4, pp 108.
Pan, Z., Tao, Z., Cao, Y.F., Wuhrer, R., and Murphy, T. (2018b). “Compressive strength and microstructure
of alkali-activated fly ash/slag binders at high temperature”. Cement and Concrete Composites, Vol.
86, pp 9-18.
Philip, P.M., Madheswaran, C.K., and Skaria, E. (2015). “Retrofitting of seismically damaged open ground
storey RCC framed building with geopolymer concrete”. Advances in Structural Engineering, Editor:
V. Matsagar, Springer, New Delhi, pp 463-481.
Phoo-ngernkham, T., Hanjitsuwan, S., Damrongwiriyanupap, N., and Chindaprasirt, P. (2017). “Effect of
sodium hydroxide and sodium silicate solutions on strengths of alkali activated high calcium fly ash
containing Portland cement”. KSCE Journal of Civil Engineering, Vol. 21, No. 6, pp 2202-2210.
Provis, J., and van Deventer, J.S.J. (2014). “Alkali Activated Materials: State-of-the-Art Report”, RILEM
TC 224-AAM. Springer.
Raj, S. D., Ganesan, N., Abraham, R., and Raju, A. (2016). “Behavior of geopolymer and conventional
concrete beam column joints under reverse cyclic loading”. Advances in Concrete Construction, Vol.
4, No. 3, pp 161-172.
Sarker, P.K. (2011). “Bond strength of reinforcing steel embedded in fly ash-based geopolymer
concrete”. Materials and Structures, Vol. 44, No. 5, pp 1021-1030.
Sarker, P.K., Kelly, S., and Yao, Z. (2014). “Effect of fire exposure on cracking, spalling and residual
strength of fly ash geopolymer concrete”. Materials & Design, Vol. 63, pp 584-592.
Sarker, P.K., and Mcbeath, S. (2015). “Fire endurance of steel reinforced fly ash geopolymer concrete
elements”. Construction and Building Materials, Vol. 90, pp 91-98.
Shaikh, F.U., and Vimonsatit, V. (2015). “Compressive strength of fly-ash-based geopolymer concrete at
elevated temperatures”. Fire and Materials, Vol. 39, No. 2, pp 174-188.
Shang, J., Dai, J.G., Zhao, T.J., Guo, S.Y., Zhang, P., and Mu, B. (2018). “Alternation of traditional cement
mortars using fly ash-based geopolymer mortars modified by slag”. Journal of Cleaner
Production, Vol. 203, pp 746-756.
Snell, C., Tempest, B., and Gentry, T. (2017). “Comparison of the thermal characteristics of Portland
cement and geopolymer cement concrete mixes”. Journal of Architectural Engineering, Vol. 23, No.
2, 04017002.
Statista (2018). World cement production by country. Online at https://www.statista.com/statistics/267364/.
Accessed on November 20, 2018.
Sumajouw, D., Hardjito, D., Wallah, S., and Rangan, B. (2007). “Fly ash-based geopolymer concrete: study
of slender reinforced columns”. Journal of Materials Science, Vol. 42, No. 9, pp 3124-3130.
Sumajouw, M., and Rangan, B.V. Low-calcium fly ash-based geopolymer concrete: reinforced beams and
columns, Research Report GC 3, Faculty of Engineering, Curtin University of Technology, Perth,
Australia, 2006.
Tao, Z., Cao, Y.F., Pan, Z., and Hassan, M.K. (2018). “Compressive behaviour of geopolymer concrete-
filled steel columns at ambient and elevated temperatures”. International Journal of High-Rise
Buildings, Vol. 7, No. 4, pp 327-342.
Tchadjie, L.N., and Ekolu, S.O. (2018). “Enhancing the reactivity of aluminosilicate materials toward
geopolymer synthesis”. Journal of Materials Science, pp 1-25.
Tchakouté, H.K., Rüscher, C.H., Kong, S., Kamseu, E., and Leonelli, C. (2016). “Geopolymer binders from
metakaolin using sodium waterglass from waste glass and rice husk ash as alternative activators: a
comparative study”. Construction and Building Materials, Vol. 114, pp 276-289.
Torres-Carrasco and Puertas, (2014). “Re-use of waste glass as alkaline activator in the preparation of
alkali-activated materials”, 34th Cement and Concrete Science Conference, 14-17 September 2014,
Sheffield, UK.
Turner, L.K., and Collins, F.G. (2013). “Carbon dioxide equivalent (CO2-e) emissions: A comparison
between geopolymer and OPC cement concrete”. Construction and Building Materials, Vol. 43, pp
125-130.
Un, C.H., Sanjayan, J.G., San Nicolas, R., and van Deventer, J.S.J. (2015). “Predictions of long-term
deflection of geopolymer concrete beams”. Construction and Building Materials, Vol. 94, pp 10-19.
van Deventer, J.S.J., Provis, J.L. and Duxson, P. (2012). "Technical and commercial progress in the
adoption of geopolymer cement". Minerals Engineering, Vol. 29, No. pp 89-104.
van Jaarsveld, J.G.S., van Deventer, J.S.J., and Lukey, G.C. (2002). “The effect of composition and
temperature on the properties of fly ash-and kaolinite-based geopolymers”. Chemical Engineering
Journal, Vol. 89, No. 1-3, pp 63-73.
Wardhono, A., Gunasekara, C., Law, D.W., and Setunge, S. (2017). “Comparison of long term performance
between alkali activated slag and fly ash geopolymer concretes”. Construction and Building
materials, Vol. 143, pp 272-279.
Xie, J., and Kayali, O. (2016). “Effect of superplasticiser on workability enhancement of Class F and Class
C fly ash-based geopolymers”. Construction and Building Materials, Vol. 122, pp 36-42.
Xu, H., and van Deventer, J.S.J. (2000). “The geopolymerisation of alumino-silicate
minerals”. International Journal of Mineral Processing, Vol. 59, No. 3, pp 247-266.
Yost, J.R., Radlinska, A., Ernst, S., Salera, M., and Martignetti, N.J. (2013). “Structural behaviour of alkali
activated fly ash concrete. Part 2: structural testing and experimental findings”. Materials and
Structures, Vol. 46, pp 449-462.
Zhang, Z., Provis, J. L., Ma, X., Reid, A., and Wang, H. (2018). “Efflorescence and subflorescence induced
microstructural and mechanical evolution in fly ash-based geopolymers”. Cement and Concrete
Composites, Vol. 92, pp 165-177.
Zhao, R., and Sanjayan, J.G. (2011). “Geopolymer and Portland cement concretes in simulated
fire”. Magazine of Concrete Research, Vol. 63, No. 3, pp 163-173.
Zuhua, Z., Xiao, Y., Huajun, Z., and Yue, C. (2009). “Role of water in the synthesis of calcined kaolin-
based geopolymer”. Applied Clay Science, Vol. 43, No. 2, pp 218-223.

View publication stats

You might also like