You are on page 1of 21

Journal of Cleaner Production 270 (2020) 122389

Contents lists available at ScienceDirect

Journal of Cleaner Production


journal homepage: www.elsevier.com/locate/jclepro

Review

Properties of fresh and hardened fly ash/slag based geopolymer


concrete: A review
Peng Zhang a, Zhen Gao a, *, Juan Wang a, **, Jinjun Guo a, Shaowei Hu a, b, Yifeng Ling c
a
School of Water Conservancy Engineering, Zhengzhou University, Zhengzhou, Henan, 450001, China
b
College of Civil Engineering, Chongqing University, Chongqing, 400045, China
c
Department of Civil, Construction and Environmental Engineering, Iowa State University, Ames, IA, 50011, United States

a r t i c l e i n f o a b s t r a c t

Article history: Geopolymers are recognized as a potentially viable alternative binder to ordinary Portland cement (OPC),
Received 14 March 2020 for reducing carbon dioxide emissions and achieving efficient waste recycling. Among the raw materials
Received in revised form of geopolymers, fly ash (FA) and ground-granulated blast-furnace slag (GGBFS) have been preferentially
15 April 2020
applied to geopolymer concrete (GPC) owing to their availability and high silica and alumina contents.
Accepted 18 May 2020
Available online 3 June 2020
The FA/GGBFS-based GPC provides a clean technology option for sustainable development. Therefore, the
specific review of FA/GGBFS-based GPC used to replace traditional concrete has become extremely
Handling editor: Baoshan Huang imperative because the related study results on the FA/GGBFS-based GPC can promote the further study
and application of this green construction material. In the present review, the reaction mechanism of
Keywords: geopolymers, as well as the properties and durability of fresh and hardened FA/GGBFS-based GPC, are
Geopolymer concrete discussed. In addition, the latest data on the FA/GGBFS-based GPC are presented. The GPC has excellent
Fly ash/slag properties and a wide range of application prospects. However, there are obstacles to its large-scale
Fresh properties applications in engineering and industry. Consequently, scholars and engineers must perform further
Mechanical properties
research to provide a complete set of theory and engineering applications of FA/GGBFS-based GPC
Durability
system.
© 2020 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Reaction mechanism of geopolymer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3. Properties of fresh GPC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.1. Workability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.2. Setting time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4. Properties of hardened GPC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
4.1. Compressive strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
4.2. Split tensile strength and flexural strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.3. Bond properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.4. Fracture behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
5. Durability of GPC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
5.1. Acid resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
5.2. Carbonation resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
5.3. Freezeethaw resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
5.4. High-temperature resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

* Corresponding author.
** Corresponding author.
E-mail addresses: zhangpeng@zzu.edu.cn (P. Zhang), gz12faafr@126.com
(Z. Gao), wangjuan@zzu.edu.cn (J. Wang), guojinjun@zzu.edu.cn (J. Guo),
hushaowei@cqu.edu.cn (S. Hu), yling@iastate.edu (Y. Ling).

https://doi.org/10.1016/j.jclepro.2020.122389
0959-6526/© 2020 Elsevier Ltd. All rights reserved.
2 P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389

5.5. Opportunities and challenges for FA/GGBFS-based GPC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17


6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Declaration of competing interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

1. Introduction activators include sodium silicate (Na2SiO3, SS) and sodium hy-
droxide (NaOH, SH). Fig. 1 shows the process flow for the produc-
Ordinary Portland cement (OPC) has always been an indis- tion of GPC. The advantage of GPC is that cement (which is used in
pensable material in the construction of large-scale hydraulic traditional concrete) is replaced with industrial waste. The GPC can
structures. It is a hydraulic cementitious material comprising a utilize industrial by-products as its raw materials and have huge
calcareous material, an appropriate amount of gypsum, and a possibility to improve the sustainability (Bajpai et al., 2020). To
prescribed mixed material. The OPCC is strong and has excellent date, FA and GGBFS have been preferentially applied to GPC owing
frost resistance. However, the production of OPC requires a large to their wide availability and high contents of silica (SiO2) and
amount of resources, incurs significant energy consumption alumina (Al2O3) (Puertas et al., 2011). Fig. 2 presents the compo-
(L€
ammlein et al., 2019; Xie et al., 2019), and leads to considerable sitions of OPC, FA and GGBFS respectively (Yang et al., 2016). As is
CO2 emissions (Zhang et al., 2018a). Peng et al. (2012) reported that shown in Table 1, the chemical composition of GGBFS is signifi-
0.66e0.82 kg of carbon emission was generated during the pro- cantly different from that of FA (Krivenko et al., 2014). Thus, the
duction of 1 kg of OPC, accounting for 5e7% of the global anthro- properties of GPC produced using different binding materials are
pogenic CO2 emissions. In the past few decades, several scholars distinct. The principal reaction products of FA/GGBFS-based GPC
have conducted extensive research on energy-efficient and sus- include calcium silicate hydrate (CeSeH gel), calcium aluminosil-
tainable building materials. Compared with OPC, geopolymers are icate hydrate (C-A-S-H gel), and sodium aluminosilicate hydrate (N-
produced using aluminosilicate-rich natural minerals and indus- A-S-H gel) (Zhuang et al., 2016).
trial waste as the primary raw materials, rather than relying on the A detailed literature review was conducted to obtain mix pro-
calcination of calcium carbonate (CaCO3). CaCO3 is the dominant portions of FA/GGBFS-based GPC and they were presented in
source of CO2 emissions in the production of OPC (Kupwade-Patil Table 2. The mix proportions can be obtained directly from the
and Allouche, 2012). Therefore, the adoption of geopolymers can literatures or calculated based on the literatures. All the experi-
reduce greenhouse-gas emissions by 73% and energy consumption mental studies listed in Table 2 have utilized the FA and the GGBFS.
by 43% (Meyer, 2009). It can be concluded that geopolymer binders It can be seen that each group is composed of different amounts of
will be a potentially viable alternative to OPC for the production of materials, and the influence of materials on the strength of GPC can
concrete (Juenger et al., 2011). be observed. The effect of all mix designs shown in Table 2 on the
Geopolymer concrete (GPC) is a promising ecofriendly building properties of FA/GGBFS-based GPC will be further analyzed in
material. Geopolymers are the products of polymerization re- subsequent sections.
actions of aluminosilicate-rich materials and alkali activators. The FA-based GPC has exhibited excellent mechanical properties
sources of these raw materials vary, e.g., fly ash (FA), ground- and durability after high-temperature curing. Generally, activation
granulated blast-furnace slag (GGBFS), and metakaolin. The of FA requires a curing temperature of 60e85  C because the
reactivity of FA is inadequate for it to be activated by alkaline ac-
tivators at the ambient temperature (Junaid et al., 2014), the

Fig. 1. Production of GPC. Fig. 2. Binder gel compositions of geopolymers and OPC.
P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389 3

Table 1
Chemical constituents: OPC, FA, and GGBFS (Yang et al., 2016; Krivenko et al., 2014).

Constituent Chemical composition (mass %)


Al2O3 Fe2O3 MnO
SiO2 Al2O3 Fe2O3 MnO MgO SiO2

OP 21.33 5.80 2.57 2.41 OP 21.33 5.80 2.57 2.41


FA 50.94 24.56 13.25 0.03 FA 50.94 24.56 13.25 0.03
GGBFS 40.0 5.91 0.32 0.50 GGBFS 40.0 5.91 0.32 0.50

Table 2
Some common mix proportions of FA/GGBFS-based GPC.

Ref. FA (Kg/m3) GGBFS (Kg/m3) Aggregates Alkaline NaOH molarity (M) Water (Kg/m3) curing temperature ( C) Strength (MPa)
(Kg/m3) activator (Kg/
m3)

Coarse Fine NaOH

Na2SiO3 Nuaklong et al. 450 1150 500 108 162 8


(2016)
24 24 40.0
450 1150 500 108 162 12 24 24 41.4
450 1150 500 108 162 16 24 24 38.4
Puertas et al. (2018) 357.0 1067.0 711.3 23.0 10 177.75 24 60.0
357.0 1276.3 563.0 42.5 10 145.0 24 62.0
Hadi et al. (2017) 0 500 1115 600 50 125 12 24 24 59.5
50 450 1115 600 50 125 12 24 24 59.0
100 400 1115 600 50 125 12 24 24 58.2
150 350 1115 600 50 125 12 24 24 49.2
200 300 1115 600 50 125 12 24 24 42.5
250 250 1115 600 50 125 12 24 24 40.9
300 200 1115 600 50 125 12 24 24 35.9
Noushini et al. (2020) 349.2 38.8 1221.2 620.8 194.0 12 13.3 23 41.7
349.2 38.8 1221.2 620.8 194.0 12 13.3 60 50.0
349.2 38.8 1221.2 620.8 194.0 12 13.3 75 62.3
349.2 38.8 1221.2 620.8 194.0 12 13.3 90 60.7
Nath and Sarker (2014) 400 0 1209 651 45.7 114.3 14 22 22 26.7
360 40 1209 651 45.7 114.3 14 22 22 34.6
320 80 1209 651 45.7 114.3 14 22 22 45.6
280 120 1209 651 45.7 114.3 14 22 22 54.8
Bernal et al. (2011) 300 990 990 111 25 25 50.0
400 885 885 148 25 25 67.8
500 780 780 185 25 25 74.4
Ding and Bai (2018) 420 1041 694 11 117 24 24 69.9
Farhan et al. (2018) 225 225 1164 627 45 112.5 14 24 24 41.1
Azzawi et al. (2018) 300 1323 623 72 108 24 24 39.0
400 1164 548 80 160 24 24 47.0
500 1005 475 100 200 24 24 62.0
Ding et al. (2018) 200 200 1068 712 12 74 24 24 50.0
300 100 1068 712 12 74 24 24 51.3
400 0 1068 712 12 74 24 24 63.4
Okoye et al. (2017) 400 1293 554 45 113 14 100 100 33.5

samples are filled with incompletely reacted gels in the process of OPCC is highly imperative. This paper summarizes the research and
reaction, which makes the polymerization cannot be reacted fully. development of FA/GGBFS-based GPC with regard to several as-
(Somna et al., 2011). It was found that GGBFS could improve the pects, such as the reaction mechanism, workability, mechanical
properties of FA-based GPC under ambient-temperature curing properties and durability. It focuses on the reaction mechanism of
conditions, possibly owing to the presence of CaO in the GGBFS geopolymers, Properties of fresh GPC, Properties of hardened GPC
(Puligilla and Mondal, 2013). Thus, the use of FA/GGBFS-based GPC and durability. It details the workability, Setting time, strength
can save energy, reduce CO2 emissions, and facilitate waste recy- properties, fracture behavior, acid resistance, carbonation resis-
cling. Additionally, FA/GGBFS-based GPC has excellent mechanical tance, freeze-thaw resistance and high-temperature resistance of
properties (Ding et al., 2016; Razak et al., 2014), freezeethaw FA/GGBFS-based GPC, and presents the problems faced by GPC
resistance (Ferdous et al., 2015), corrosion resistance (Ariffin technology.
et al., 2013), high-temperature resistance (Khater, 2014), and
outstanding interfacial bond properties and other unique charac-
2. Reaction mechanism of geopolymer
teristics (Moon et al., 2014).
The FA/GGBFS-based GPC has been studied by many scholars
A geopolymer is a three-dimensional zeolite-like network
and engineers, which poses the excellent properties. However,
structure comprising silicon-oxygen (SieO) and aluminum-oxygen
there is no the review of the properties and durability of FA/GGBFS-
(AleO) tetrahedral (Davidovits and Cordi, 1979; Davidovits, 1989).
based GPC. In order to better sum up the relevant content to pro-
Several studies have been performed on the reaction mechanism of
mote the application and further research of FA/GGBFS-based GPC,
geopolymers via the dissolution of the silicon and aluminum ma-
the specific review of the FA/GGBFS- based GPC used to replace the
terials in a highly alkaline environment (Li et al., 2019). The reaction
4 P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389

mechanism of geopolymer is related to the preparation technology,


the amount of alkali activator and the chemical composition of raw
materials (Rees et al., 2008). Because of the multiple components in
the liquid phase, all the reactions occur simultaneously and affect
each other (Zhang et al., 2011). Therefore, the geological polymer-
ization reaction processes are more complex than the hydration of
OPC. Davidovits suggested that the structure of a geopolymer is as
follows:

Mn ½ðSiO2 ÞZ  AlO2 n ,wH2 O (1)

where, M is an alkali metal cation; n represents the degree of


polycondensation; w represents the number of chemically bound
water molecules; z represents the silicon-to-aluminum ratio (Si/Al), Fig. 4. Conceptual model for geopolymerization (Glukhovsky, 2011).
which is 1, 2, or 3. The structures of the geopolymer are divided into
three types according to the Z value: single silicon-aluminum (Gel
1, Si/Al ¼ 1), double silicon-aluminum (Gel 2, Si/Al ¼ 2), and triple three-dimensional network structure of the geopolymer.
silicon-aluminum (Gel 3, Si/Al ¼ 3). These structures are illustrated The role and existence form of water in geopolymer are related
in Fig. 3. to the way of geopolymerization. Qualitative analysis indicates that
The hardening of geopolymer materials involves breaking and water is released in the process of reaction (Hardjito et al., 2005).
recombining the SieO and AleO bonds under the action of an The evaporated water left discontinuous nanopores, improving the
alkaline activator. The model was pioneered by Davidovits (1991). performance of the geopolymer. Thus, water plays no role in the
The reaction process mainly involves the following four steps. (ⅰ) chemical reaction processes, and it simply enhances the work-
Dissolution: The SieO and AleO bonds are broken in the alumi- ability of the mixture during production. This differs from the hy-
nosilicate materials under the action of the alkali activator, and the dration of OPCC, where water is involved in a chemical reaction. In
SieO and AleO tetrahedral monomers are released. (ⅱ) Diffusion: addition to the aforementioned models, many scholars have
The dissolved SieO and AleO tetrahedral monomers diffuse into investigated the reaction mechanism of geopolymers, such as
the reaction system. According to the principle of chemical equi- alkaline activation (Fernandez et al., 2006, 2008), SS/SH ratio(-
librium, the silicon and aluminum concentrations decrease on the Duxson et al., 2005) alkali cations on aluminum (Liew et al., 2016)
particle surface owing to the diffusion, and the dissolution process and interfacial bond mechanism between fibers and geopolymer
continues. (ⅲ) Polycondensation: SieO and AleO tetrahedra form composite (Ranjbar et al., 2016), which propels the geopolymer
amorphous -Si-O-Al-O- structures or zeolite crystals through models more reasonable and complete.
polymerization. (ⅳ) Hardening: a dehydration reaction occurs,
forming a hardened geopolymer with high mechanical strength. 3. Properties of fresh GPC
These steps are performed simultaneously. However, only the
leading step is different during the reaction process. For geo- 3.1. Workability
polymer materials with different raw materials, the polymerization
mechanisms are not exactly the same, but they essentially conform An alkaline solution (NaOH, Na2SiO3) is used as a liquid
to the foregoing reaction process. component in GPC. Owing to the high viscosity of waterglass
According to the conceptual model (Glukhovsky, 2011), acti- (Na2SiO3 solution, WG), the mixture is sticky (Krivenko et al., 2014;
vated silico-aluminous cementitious materials form geopolymers Li et al., 2018). Therefore, the rheology of GPC differs from that of
under the action of alkaline activators. The model divides the OPCC.
polymerization reactions into deconstruction, gel formation, poly- The influence of each component on the workability of FA/
condensation, and crystallization. As is shown in Fig. 4, there are GGBFS-based GPC must be analyzed. The effects of alkaline solu-
five types of reaction processes for the geopolymer in the Glu- tions with concentrations of 35%, 40%, and 45% on the performance
khovsky model. In contrast to the Davidovits model, the poly- of GGBFS-based GPC were analyzed, the results indicated that the
condensation process is divided into the two processes of gelation slump of the GPC was maximized when the alkaline-solution
and reconstitution. Gelation involves dehydration to produce an concentration was 45% (Nath and Sarker, 2014). However,
oligomeric gel when the aluminosilicate is in a saturated state, and increasing the SH concentration reduced the slump of the GPC
reconstitution involves the rearrangement of the oligomeric gels. (Nuaklong et al., 2016), possibly because it led to the penetration of
The gels are dehydrated and condensed, gradually forming the more SiO2 and Al2O3, accelerating the geopolymer reaction

Fig. 3. Geopolymer structures according to Davidovits (1989).


P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389 5

processes and improving the stiffness of the system.


The SS/SH ratio significantly affected the workability of FA-
based GPC, and different SS/SH ratios can cause unstable varia-
tions in the workability of GPC (Nath and Sarker, 2015). The vis-
cosity of the GPC increased with SS/SH. Generally, the slump of the
GPC decreased with an increase in the Na2SiO3 content. Hardjito
(2005) indicated that H2O/Na2O was suitable for FA-based GPC in
the range of 10e14. This was confirmed by Prasittisopin and
Sereewatthanawut (2018). Additionally, the slump increased
owing to the free water in the superplasticizer. This increase can be
expressed by a linear regression of the data, which yields the
follows:

slump ¼ 2112ðw : bÞ þ 1275ðsp : bÞ  61 (2)

where,w : b represents the water-to-binder ratio, and sp : b repre-


sents the superplasticizer-to-binder ratio.
Puertas et al. (2018) used two mixing protocols to prepare OPCC
and GGBFS-based GPC (Fig. 5). Protocol 1 principally considered the
Fig. 6. Slump values for OPCC and GPC (Puertas et al., 2018).
stability and mechanical properties of the concrete according to
RILEM TC 247-DTA, and Protocol 2 had a longer mixing time, ac-
cording to previous rheological studies of geopolymer concrete Rao, 2017). This may have been due to the high calcium content
(Tuinukuafe et al., 2019) and mortar (Puertas et al., 2014; Alonso in the GGBFS, which accelerated the reaction process of the geo-
et al., 2017). The slump values of OPCC and GPC with WG acti- polymers by forming an amorphous CaeAleSi gel. This was
vated and NaOH activated (Fig. 6) indicate that the slump of GPC confirmed by Cheah et al. (2017).
(particularly with WG activated) was greater than that of OPCC. It
was also observed that the mixing time affected the slump, and
longer periods reduced the workability of the concrete. In partic- 3.2. Setting time
ular, the workability of the GPC exhibited a large decline, and it
could not be measured after 30 min. Protocol 2 had a beneficial The setting time of GPC is a vital technical parameter in the
effect on the concrete and maintained the workability. process of construction (Won et al., 2009), It is divided into initial
For GPC mixed with FA and GGBFS cementitious materials, some setting time and final setting time, which play a significant role in
results revealed that replacing part of the FA with GGBFS reduced construction. The initial setting time should not be too short, and
the slump value of the GPC (Laskar and Talukdar, 2017; Venu and the final setting time should not be too long (Embong et al., 2016).

Fig. 5. Mixing protocols for OPCC and GPC (Puertas et al., 2018).
6 P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389

However, FA/GGBFS-based GPC has the characteristic of rapid different thermal curing methods. Fig. 8 showed the 28-
hardening; thus, it is necessary to investigate the relationship be- d compressive strength of GPC samples at ambient temperature
tween the setting time and the influencing factors. A mechanism and after thermal curing. The curing temperature significantly
analysis is also necessary. affected the compressive strength of the FA-based GPC. In the cured
As is shown in Fig. 7, the initial and final setting time of FA-based samples, the initial growth rate of the compressive strength
GPC both decreased with the increasing GGBFS and OPC contents, increased with the curing temperature and duration, and the
and they increased when the content of the alkali activator optimal compressive strength was obtained at 75  C (Noushini
increased. With the other mixing variables unchanged, increasing et al., 2020). This may be because higher curing temperatures and
the SS/SH ratio can reduce the setting time (Nath and Sarker, 2014, longer durations led to the formation of more reaction products;
2015). Hanjitsuwan et al. (2014) investigated the effect of the SH thus, with the proper curing temperature, more metal ions were
concentration on GPC. As is shown in Table 3, the initial and final incorporated into the geopolymer matrix, and the concentration of
setting times of the GPC increased with the SH concentration. This metal ions decreased. Noushini et al. investigated the leaching of
is because the leaching of SiO2 and Al2O3 is slow at a low SH con- alkali metals by measuring the pH values of hardened geopolymer
centration (Rattanasak and Chindaprasirt, 2009), and the leaching paste, as is shown in Table 4. The results indicated that the pH
of Ca2þ into the solution continues; thus, the solution contains a values of geopolymers subjected to thermal curing were lower than
large amount of calcium. The amount of calcium is sufficient to those of geopolymers subjected to ambient-temperature curing.
precipitate and participate in the hydration reaction to form a The excessive alkalinity in the case of ambient-temperature curing
CeSeH gel, reducing the setting time of the mixture (Alonso and suggests a lower degree of polymerization of the samples, as well as
Palomo, 2001; Chindaprasirt et al., 2012). reduced formation of the reaction product. Therefore, the
The setting times of GPC samples with different proportions of compressive strength of the GPC subjected to ambient-temperature
FA (replacing GGBFS) were examined by Hadi et al. (2017). The curing was lower (Zhang et al., 2014).
proportion of GGBFS replaced with FA ranged from 10% to 60%. As is FA-based GPC exhibited the optimal compressive strength with
shown in Table 3, the setting time of the GPC increased with the ambient-temperature curing when a suitable mix proportion was
addition of FA, which is explained as follows. (ⅰ) FA particles used (Nath and Sarker, 2014). The effects of different alkaline-
improve the fluidity of each component of the mixture owing to solution contents and the SS/SH ratio on GPC samples are pre-
their smooth spherical shape, and the setting time of GPC is sented in Fig. 9. The molar ratios and water-to-solid ratios (w/s) of
extended. (ii) FA cannot form geopolymers immediately at ambient the mixtures are presented in Table 5. The Na2O/SiO2 ratio
temperature; thus, a longer setting time is required within a increased as the SS/SH ratio decreased. A higher Na2O/SiO2 ratio
reasonable range (Nath and Sarker, 2015; Go €rhan and Kürklü, corresponded to a higher compressive strength of the samples. This
2014).
There are many factors affecting the workability and setting
time of GPC. Due to different raw materials and curing conditions, Table 3
the optimal mix proportion proposed varies (Kastiukas et al., 2020). Effect of NaOH concentration and FA content on the setting time (Hanjitsuwan et al.,
Moreover, the workability of the prepared GPC fluctuates greatly. 2014; Hadi et al., 2017).
Therefore, the preparation of GPC with relatively stable perfor- Content Initial setting time Final setting time
mance is the key problem to be solved in the application of GPC. (min) (min)

NaOH concentration 8 30 130


4. Properties of hardened GPC (molar) 10 55 160
12 75 202
4.1. Compressive strength 15 100 238
FA/(FA þ GGBFS) (%) 0 25 56
10 30 66
The compressive strength significantly affects the structural 20 35 68
stability and safety. The compressive strength of FA/GGBFS-based 30 45 78
GPC is affected by the curing conditions and the raw materials 40 60 83
(Hansen et al., 2018; Le et al., 2017). 50 66 92
60 76 105
Noushini et al. prepared low-calcium FA-based GPC using 12

Fig. 7. Effects of different phases on the setting time of FA-based GPC (Nath and Sarker, 2014, 2015).
P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389 7

Fig. 8. Compressive strength of ambient and heat-cured samples (Noushini et al., 2020).

Table 4 polymerization process and increased the amount of oligomeriza-


PH values of hardened geopolymer paste (Noushini et al., 2020). tion reaction products (Ruiz et al., 2012). Although the Si/Al to the
Curing pH of values of leaching solutions total binder increased when the alkali-activator content increased,
3d 7d 28 d
the w/s was the key factor affecting the compressive strength. The
results indicated that the alkaline-solution content in the geo-
OPC-23C 12.71 12.72 12.72
polymer mixture was a significant parameter for ensuring adequate
GPC-23C 12.64 12.5 12.29
GPC-60C-8h 12.29 12.27 12.15 compressive strength without affecting the workability and setting
GPC-60C-12h 12.2 12.18 12.14 time.
GPC-60C-18h 12.17 12.14 12.12 Prasittisopin and Sereewatthanawut (2018) examined the effect
GPC-60C-24h 12.16 12.14 11.98
of sp : b on the compressive strength of OPCC and FA-based GPC.
GPC-75C-8h 12.15 12.1 11.98
GPC-75C-12h 12.1 12.06 11.96
The results indicated that the superplasticizer could increase the
GPC-75C-18h 12.08 12.02 11.91 compressive strength of the OPCC; however, the compressive
GPC-75C-24 12.07 12 11.9 strength of the GPC decreased with an increase in sp : b. Similar
GPC-90C-8h 12.15 12.08 11.98 results were reported by AlBakri et al. (2012). By referring to pre-
GPC-90C-12h 12.09 12.05 11.96
vious studies, researchers can avoid the impact of superplasticizers
GPC-90C-18h 12.05 12.02 11.9
GPC-90C-24h 12.04 12.02 11.88 on GPC when conducting related experiments in the future.
Therefore, these works contribute to prospective experiments.
Bernal evaluated the compressive strength of GPC with different
slag contents and used the compressive strength of OPCC as a
reference for comparison. The compressive strength of GPC and
OPCC samples with different slag contents are presented in Fig. 10.
The strength increased with the binder content (slag for GPC,
cement for OPCC). Regardless of the binder content, the GPC had a
higher compressive strength than the OPCC (Bernal et al., 2011).
Therefore, the GPC samples had excellent mechanical properties.
The polymerization of GPC was controlled by the dissolution and
precipitation processes. When the pH value was increased, the slag
diffusion-controlled reaction processes of GPC were faster than the
hydration reaction of OPC, and the GPC had a higher compressive
strength in the early curing period. The hydration products for the
OPCC were primarily calcium hydroxide (Ca(OH)2), CeSeH gel, and
ettringite (AFt), and those for the GGBFS-based GPC included geo-
polymer gel and CeSeH gel (Fig. 11) (Chen et al., 2019). Comparing
Fig. 11(a) and (b) reveal that the structure of the GGBFS-based GPC
was denser than that of the OPCC, which explains the higher
strength of the GPC. Fig. 11(c) and (d) showed scanning electron
Fig. 9. Effect of the alkaline activator content (A) and Na2SiO3/NaOH ratio (R) on the
microscopy (SEM) images of the interfacial transition zones (ITZ) in
compressive strength of the samples (Nath and Sarker, 2014). the OPCC and GPC samples. The ITZ of GPC was highly dense and
uniform, and the particle roughness enhanced the mechanical
integrity of GPC (Pasupathy et al., 2017; Mehta and Siddique, 2017a)
may have been due to the increase in the amount of Na2O in the and recycled GPC (Shi et al., 2012). Notably, the microstructure of
mixture, which enhanced the mechanical properties of the geo- the GPC differed from that of the OPCC, and the ITZ was the weakest
polymers (Chi and Huang, 2013). The strength decreased as the part of the OPCC and had a high porosity. This explains why the ITZ
alkaline-solution content increased, owing to the high w/s of the of GPC had higher bond strength than that of OPCC (Lee and Van,
mixture with a high liquid content, which hindered the 2007).
8 P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389

Table 5
Criticalmolar ratios and w/s values for the samples (Nath and Sarker, 2014).

Designation Alkaline activator content (%) Na2SiO3/NaOH Molar ratio Water/solid

Na2O/SiO2 H2O/Na2O Si/Al

A35 35 2.5 0.11 11.67 1.77 0.18


A40/R2.5 40 2.5 0.12 11.75 1.80 0.2
A45 45 2.5 0.13 11.80 1.84 0.22
R1.5 40 1.5 0.14 10.63 1.76 0.20
R2.0 40 2.0 0.13 11.25 1.78 0.20

GPC was primarily due to the ability of the fibers to transmit stress
and load. Nevertheless, increasing the fiber content cannot
continuously enhance the compressive strength of GPC (Ding and
Bai, 2018). This is consistent with previous studies on steel fiber-
reinforced concrete (Song and Hwang, 2004) and may be due to
excessive fibers making the mixing difficult, which adversely af-
fects the workability and uniformity.
Many studies have been performed to enhance the mechanical
properties of GPC. The incorporation of nano-silica (NS) improved
the compressive strength of GPC (Ibrahim et al., 2018), as is shown
in Fig. 13. The compressive strength of the GPC was maximized with
the incorporation of 5% NS. This was explained as follows: (1) the
NS had a strong pozzolanic activity, it enhanced the conversion of
raw materials into CeSeH gel and C(N)-A-S-H gel; (2) the NS par-
ticles had filling and nucleation effects. However, the incorporation
of 7.5% NS reduced the compressive strength of GPC, possibly owing
to the accumulation of nanoparticles in the geopolymer structure
(Yip and Deventer, 2003; Yip et al., 2005).
Fig. 10. Effect of the slag content on the compressive strength for GPC and OPCC
(Bernal et al., 2011).
4.2. Split tensile strength and flexural strength

As is shown in Fig. 12, basalt fiber significantly increased the The split tensile strength and flexural strength of concrete are
compressive strength of GPC. With the increasing fiber content, the important mechanical properties. The split tensile strength is
compressive strength first increased and then decreased. The related to the initiation and expansion of cracks in the concrete
optimal basalt fiber content was 2.0%. Additionally, there was little structure, the shearing and anchoring of bars, and other phenom-
difference in the compressive strength of the GPC between the ena. The flexural strength indicates the ability of beams to resist
curing durations of 7 and 28 d, in contrast to the case of OPCC. The bending and destruction.
compressive strength of OPCC increased with the curing duration The effects of the GGBFS content on the split tensile strength
(Ronad et al., 2016). The increase in the compressive strength of the and flexural strength of FA-based GPC were examined by Fang et al.
(2018). As is shown in Fig. 14, a higher GGBFS content yielded a

Fig. 11. SEM images of the binders and ITZs of OPCC and GPC, after 28 d (Chen et al., Fig. 12. Compressive strength of GPCs with the incorporation of basalt fiber at 7 and
2019). 28 d (Ronad et al., 2016).
P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389 9

Fig. 13. Effect of the NS content on the compressive strength of samples (Ibrahim et al., Fig. 15. Split tensile strength and flexural strength of GPC with respect to the slag
2018). content and molarity at 28 d (Singhal et al., 2018).

higher split tensile strength and flexural strength of the GPC. This crystallinity, and the strength was higher (Mehta and Siddique,
was attributed to the high activity of GGBFS, which promoted the 2017b).
formation of C-A-S-H gel and N-A-S-H gel, accelerating the reaction Wongsa et al. tested the split tensile strength of FA-based GPC
process (Kumar et al., 2010). using SS/SH ratios of 0.5, 1.0, and 1.5 and alkaline solution-to-FA
Fig. 15 shows the effects of the SH content and slag content on ratios of 0.70, 0.75, and 0.80. The results are presented in Fig. 17.
the 28-d strength of GPC cured at 27  C (Singhal et al., 2018). The The split tensile strength of the GPC decreased with the increasing
trend of the flexural strength of the specimens is consistent with SS/SH ratio, and the GPC exhibited excellent mechanical properties
that of the split tensile strength. The strength increased with the SH when the alkali activator-to-FA ratio was 0.7 (Wongsa et al., 2016).
concentration and slag content. The X-ray diffraction (XRD) analysis In general, the different types of strengths of concrete are closely
results for FA-based GPC cured for 28 d at ambient temperature are related. With regard to OPCC, various codes and standards are
presented in Fig. 16. Detailed information regarding the crystal available for predicting the split tensile strength (Bhanja and
phase was obtained. In addition to sharp diffraction peaks, amor- Sengupta, 2005) and flexural strength according to the compres-
phous states were detected. For all the samples, diffraction peaks of sive strength. These relationships are based on extensive data
quartz were clearly observed. All the other reflections can be clearly regarding different engineering properties, which have been
observed in Fig. 16, corresponding to mullites, merwinites, and collected for decades (ACI Committee 318, 2011; Darwin et al.,
CeSeH gels. The proportion of amorphous silicate compounds was
high in the GPC prepared with an 8 mol. % SH solution. Thus, with a
higher concentration of the alkaline solution, the GPC had a higher

Fig. 14. Split tensile strength and flexural strength of GPC with respect to the slag Fig. 16. XRD patterns of GPC specimens with different molarities at 28 d (Singhal et al.,
content at 28 d (Fang et al., 2018). 2018).
10 P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389

2016). However, the polymerization process of GPC differs from the


hydration reaction process of traditional concrete. Many scholars
have investigated compressive strength of GPC, split tensile
strength (Sofi et al., 2007), flexural strength (Diaz et al., 2011), and
the relationships among the strengths of GPC cured at ambient
temperature (Lee and Lee, 2013; Phoongernkham et al., 2014) and
calculation formulas of mechanical parameters have been proposed
(Nath and Sarker, 2017), as is shown in Figs. 18 and 19. These for-
mulas are useful for further research.

4.3. Bond properties

The bearing capacity and performance of reinforced concrete


depend on the strength of the bond between the concrete and the
bars. The interaction between the concrete and bars is composed of
three independent parts: chemical adhesion, friction, and me-
chanical occlusion (Dahou et al., 2016). The bond strength of GPC is
affected by many factors, and scholars have evaluated the bonding
properties of reinforced GPC. The behavior of the bond stress to
free-end slip is divided into three stages, as is shown in Fig. 20
Fig. 18. Correlation between the split tensile strength ðfst Þ and the compressive
(Farhan et al., 2018). In the first stage, the bond stresseslip strength ðfc Þ of GPC at 28 d.
response increased linearly. In the second stage, the bond
stresseslip response was highly nonlinear. In the final stage, the
specimens reached the maximum bonding stress, and longitudinal
cracks appeared.
Fig. 21 presents the bond strength of GPC to bars with different
SS/SH ratios and FA contents. As the SS/SH ratio increased from 1.5
to 2.5, the GPC bond strength increased. Additionally, with the
increasing FA content, the bond strength between the GPC and the
bars increased (Azzawi et al., 2018).
The bonding between glass fiber-reinforced polymer bars and
GPC was evaluated via a direct pullout test. The relationship be-
tween the average bonding stress of GPC and the bar diameter is
shown in Fig. 22. The failure of the specimens was due to the
pullout of the bars, and an increase in the diameter of bars reduced
the peak bond stress (Maranan et al., 2015).
Adak examined the effects of different curing conditions and
types of reinforcement on the bonding properties of low-calcium
FA-based modified GPC (Adak and Mandal, 2019). The bond
strengths of GPC and OPCC with deformed steel bars and mild steel
bars after 28-day presented in Fig. 23. Here, GPC I represents the

Fig. 19. Correlation between the flexural strength ðft Þ and the compressive strength
ðfc Þ of GPC at 28 d.

Fig. 17. Split tensile strength of GPC with different SS/SH and binder/FA ratios at 28 d
(Wongsa et al., 2016). Fig. 20. General behavior of bond stress versus slip (Farhan et al., 2018).
P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389 11

4.4. Fracture behavior

The fracture behavior of concrete is affected by the material


properties, such as the mixing composition, and type of aggregate.
For geopolymers to be used as substitutes for cement, their prop-
erties should be extensively studied to ensure their applicability in
structures where the fracture behaviors are crucial (Pan et al., 2011;
Wang et al., 2019).
Rao et al. examined the fracture behavior of GPC and performed
a pre-notched three-point bending test with notch depth/beam
depth (a0/h) ratios of 0.1, 0.15, and 0.2 and different curing condi-
tions (in direct sunlight and in an oven at 60  C for 24 h) (Rao et al.,
2015). The size parameters of the samples are presented in Fig. 24,
and the experimental results are shown in Fig. 25. Compared with
GPC cured at the ambient temperature, the GPC samples cured at
60  C for 24 h exhibited better structural properties with regard to
the fracture parameters. The results indicated that the energy of the
GPC cured at 60  C was nearly 50% higher than that of the GPC
cured at the ambient temperature.
The effects of the material parameters, such as the alkali-
Fig. 21. Bond strength of GPC with different FA contents and SS/SH ratios (Azzawi solution concentration, alkali-activator modulus, and FA/GGBFS
et al., 2018). and liquid-to-binder ratios, on the fracture performance of FA/
GGBFS-based GPC were investigated by Ding et al. (2018). The
complete loadedisplacement (P-D) curve reflected the failure
behavior of the GPC. The average P-D curve for the GPC beams is
thermal activation of the FA and the alkaline solution for 45 min at shown in Fig. 26.
60  C before casting, and GPC II represents the thermal curing of the The peak load and fracture energy of the beams were calculated,
specimens at 60  C for 48 h after casting. The specimens were kept and the fracture parameters are presented in Fig. 27. The peak load
at the ambient temperature prior to the testing. The results indi- and fracture energy increased as the alkali concentration and slag/
cated that the bond strengths for GPC I and II were higher than that FA ratio increased, and the trends were similar, as is shown in
of the OPCC, for both deformed steel bars and mild steel bars. GPC I Fig. 27(a) and (c). As the modulus of the alkali activator increased,
exhibited superior bonding behavior to GPC II. The failure of the the peak load of the beams increased slightly, and the fracture
samples was caused by the pullout of the steel bars. As expected, energy increased significantly. The fracture energy of the beams
during the loading process, the slip of the mild steel bars was increased with an increase in the alkali-activator modulus. The
greater than that of the deformed steel bars. The mechanical beams exhibited excellent ductile behavior at Ms ¼ 2.0, as is shown
strength and durability of GPC I were better than those of the in Fig. 27(b). However, the peak load and fracture energy of the
traditional heat-cured GPC (GPC II) and OPCC. The early thermal beams decreased as the liquid-to-binder ratio increased, as is
activation of GPC I accelerated the homogeneous and smooth shown in Fig. 27(d).
polymerization of FA, facilitating the amorphous-to-crystalline Ding and Bai (2018) also evaluated the average fracture energy
transition of FA in the geopolymer matrix. of steel fiber-reinforced GPC beams, as is shown in Table 6. The
energy-absorption capacity of the GPC beams was significantly
improved with the incorporation of the fibers. The steel fibers were
pulled out of the matrix, which was consistent with a previous
study (Lin et al., 2008). The fibers have the function of cracking and
toughening, and the geopolymer binder has a uniform micro-
structure and excellent binding performance with fibers (Castel and
Foster, 2015), the incorporation of fibers can significantly increase
the fracture energy of GPC (Aydın and Baradan, 2013).
Geopolymer has the advantages of rapid hardening and excel-
lent strength, however, its high shrinkage rate limits further
development (Qian et al., 2020). Therefore, it is urgent to find a way
to improve the shrinkage. Otherwise, it is the focus of future work
to study the basic mechanical property index of GPC similar to
OPCC, establish the full stress-strain curve equation, and adjust the
shrinkage and creep model applicable to GPC.

5. Durability of GPC

5.1. Acid resistance

In general, the damage to the cement binder under acid attack is


mainly caused by the reaction of calcium compounds with the acid
solution, which generates tensile stress and induces cracking of the
Fig. 22. Average bond stress versus bar diameter for GPC specimens (Maranan et al., cement binder system. However, FA/GGBFS-based GPC has excel-
2015). lent acid resistance. This is attributed to the low water absorption
12 P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389

Fig. 23. Bond stresseslip curves of GPC and OPCC with and (Adak and Mandal, 2019).

Fig. 24. Schematic of the specimen used for the three-point bending test.

Fig. 26. P-D curves of slag/FA-based GPC beams (Ding et al., 2018).

However, the weight loss of the GGBFS-based GPC was <10% after it
was exposed to the H2SO4 solution for 365 d. The corresponding
loss of the compressive strength is shown in Fig. 28(b) (Fernando
et al., 2010). The compressive-strength loss of the GPC after 365 d
of exposure to the H2SO4 solution (36.4%e39.1%) was significantly
smaller than that of the OPCC (almost 90%) after 90 d of exposure.
These results indicate that all the GPC samples had better corrosion
resistance than the OPCC. The excellent acid resistance of GPC may
have been due to the decalcification of the C-A-S-H gel in the ma-
trix. The significant reductions of the weight loss and compressive-
strength loss may have been due to the dense microstructure and
splendid pore structure, which resulted in low penetration of the
acid solution, providing less opportunity for further reactions
(Bakharev et al., 2003).
Fig. 29 showed the effect of a 2% H2SO4 solution on the dura-
bility of FA-based GPC and OPCC over 90 d (Okoye et al., 2017). The
appearance of the sample indicated that the GPC containing 20%
Fig. 25. Fracture test results for different types of concrete (Rao et al., 2015). SiO2fume in the acid solution was not deteriorated and that the
surface was not corroded. However, the OPCC exhibited surface
corrosion and edge breakage. Among the samples tested, the OPCC
and calcium content of the geopolymer, which result in less soluble exhibited the largest weight loss and compressive-strength loss in
compounds (Fernando and Said, 2011). the H2SO4 solution. The GPC with 20% silicon fume instead of FA
The long-term durability of GGBFS-based GPC in a corrosive exhibited the smallest weight loss and compressive-strength loss,
environment (5% sulfuric acid (H2SO4) solution) was examined by as is shown in Fig. 30. The poor performance of the OPCC was due to
Manjunath et al. (2019). As is shown in Fig. 28(a), the weight loss of the reactions of the Ca(OH)2 and CeSeH gels with the acid solution.
OPCC was approximately 75% at 56 d, and the degradation loss was The OPCC was susceptible to acid attack because of the CeSeH gel.
100% after the OPCC was exposed to an H2SO4 solution for 103 d. The formation of the N-A-S-H and C-A-S-H gels rendered the GPC
P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389 13

Fig. 27. Effects of different parameters on the peak load and fracture energy of GPC (Ding et al., 2018).

Table 6 highly resistant to acid attack. The FA and SiO2 fume contained a
Fracture energy for plain and steel fiber-reinforced GPC (Ding and Bai, 2018). large amount of aluminosilicate in the GPC, which reacted with
GPC-SF0 GPC-SF1.0 GPC-SF1.5 GPC-SF2.0 Ca(OH)2 to produce a stable C-A-S-H gel and filled the pores to a
Fracture energy, GF (N/m) 207.4 4884.5 6626.3 6203.8
considerable extent, reducing the weight loss (Aydın et al., 2007).
203 3717.3 4874 5037.4 The effects of an H2SO4 solution on the microstructures of FA-
213.6 4102 5626.6 5048 based GPC and OPCC were investigated by Gu et al. (2018). The
207.4 3996.8 6373.9 5731.3 ITZ of the OPCC before the acid attack exhibited close contact be-
Average GF (N/m) 207.9 4175.1 5875.2 5505.1
tween the aggregate and the matrix (Fig. 31(a)). Gypsum crystals
K 1 20.1 28.2 26.5
were formed near the ITZ after 142 d of erosion in the H2SO4
Note: The steel fiber contents were 1.0, 1.5, and 2.0 vol%.

Fig. 28. Weight loss and compressive strength of GPC and OPC exposed to 5% H2SO4 for different durations Manjunath et al. (2019).
14 P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389

Fig. 29. Visual appearance of four blocks of samples exposed to 5% H2SO4 for 90 d (Okoye et al., 2017).

solution (Fig. 31(b)). The porous properties of gypsum reduced the 5.2. Carbonation resistance
strength of the ITZ, which was consistent with a previous report
(Alexander et al., 2013). The ITZ of the GPC samples before the acid The carbonation of OPCC is a chemical reaction between
attack exhibited a tight connection. As is shown in Fig. 32, the Ca(OH)2 and CO2 penetrating the interior of concrete, which results
morphologies of the ITZs of GPC samples before and after immer- in the production of CaCO3. The penetration of CO2 reduces the PH
sion in the H2SO4 solution were not significantly different, indi- of the concrete, which reduces the carbonation and eliminates the
cating that the ITZ in the GPC had better acid resistance than that in protective alkaline layer around the steel bars. This is the main
the OPCC. Provis and van Deventer (2009) observed a similar reason for the corrosion of steel bars in concrete. The carbonation
phenomenon. mechanism of GPC differs from that of traditional concrete. For
Additionally, through XRD analysis, Gu found that the chemical high-calcium GPC, owing to the lack of Ca(OH)2 in the polymeri-
composition of the OPCC was significantly changed after the acid zation products, the CO2 in the GPC directly reacts with the C-A-S-H
attack. However, the degradation process of the GPC was slow in gel, producing CaCO3. For low-calcium GPC, because the primary
the initial stage. Crystalline gypsum was detected in the samples product is a N-A-S-H gel without a decalcification process, the
attacked in the H2SO4 solution for 14 weeks, which retained most of primary process of carbonation is the transformation of the pore
their original properties after the 142 d of acid attack. Owing to the solution from high-alkalinity to high-Na2CO3 concentration, and
extremely low calcium content in class F FA, the gypsum was the microstructure of the matrix changes slightly.
possibly formed by the reaction between CaCO3 in the coarse Bernal (2011) and Behfarnia (2017) examined the factors
aggregate of limestone and H2SO4 solution. affecting the carbonation of GPC (Table 7 and Fig. 33). As the
alkaline solution/slag (AS/S) ratio increased, the alkalinity of the

Fig. 30. Weight loss and compressive-strength loss of GPC and OPCC exposed to 2% H2SO4 for different durations (Okoye et al., 2017).
P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389 15

Fig. 31. ITZ change of OPCC samples Gu et al. (2018).

Fig. 32. ITZ change of GPC samples Gu et al. (2018).

solution increased, which caused the alkalinity of the GPC to the different chemical compositions of GGBFS and OPC.
decrease slowly and reduced the carbonation depth. This is similar
to the findings of Bernal (2015). However, only the slag content was 5.3. Freezeethaw resistance
increased. As is shown in Fig. 33, both GPC and OPCC exhibited a
smaller carbonation depth with an increase in the slag content, but Damage due to freezing significantly affects the performance of
the carbonation depth of OPCC was smaller than that of the GPC GPC, posing a threat to the durability and safety of GPC structures.
(Bernal et al., 2011). During the exposure of GPC samples to CO2, the Pilehvar et al. (2019) examined the effect of the freezing conditions
principal reaction product was the carbonate compound, which on the properties of OPCC and FA/GGBFS-based GPC containing
was produced by carbonation of the alkaline pore solution and the microcapsule phase-change materials (MPCMs), as is shown in
C-A-S-H gel. Fig. 35. The results indicated that the GPC had a better freeze
Fig. 34 showed the effects of water curing and plastic-cover resistance than the OPCC. With the addition of MPCMs, the strength
curing on the carbonation depth of GGBFS-based GPC. Carboniza- of all the samples decreased to <2.5% after 28 d of freezeethaw
tion performance of water-cured samples was more excellent than cycles. Thus, the MPCM had distinct frost resistance, which is
the plastic cover curing. This may be because the chemical reaction attributed to the microcapsules providing free space when the
of the binder required water, and the gel was not formed uniformly water froze, which reduced the stress caused by frost.
in the matrix when water was scarce, resulting in a low perme- Cai et al. (2017) used fractal theory to investigate the fracture of
ability resistance of the CO2. However, increasing the proportion of GGBFS-based GPC during freezeethaw cycling. The freezeethaw
micro-silica can significantly improve the carbonization perfor- damage area was calculated using the crack distribution. As is
mance. Micro-silica had the advantages of small size, aggregate shown in Fig. 36, the fractal dimension had an excellent linear
function and volcanic ash properties, the replacement of the GGBFS relationship with the fracture toughness. Therefore, the fracture
with micro-silica had a positive effect on the carbonation depth of toughening of the GPC was reflected by the fractal dimension. The
GPC samples (Leemann et al., 2015). The difference in the effects of experimental results in Fig. 37 indicate that with increasing
micro-silica on the performance of GPC and OPCC may be caused by freezeethaw cycles, the fractal dimension increased, and the
16 P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389

Table 7
Carbonation-depth values for GPC and OPCC (mm).

Mix Carbonation depth (14 d) Carbonation depth (28 d) dGPC/dOPCC (14 d) dGPC/dOPCC (28 d)

OPCC 4.93 6.12 1 1


S40 7.34 8.93 1.49 1.46
S45 7.09 8.71 1.44 1.42
S50 6.87 8.13 1.39 1.33
S55 6.8 7.78 1.38 1.27

Note: Ratio of S40 to AS/S ¼ 0.4.

5.4. High-temperature resistance

The geopolymer is a three-dimensional zeolite-like network


structure that requires a higher temperature than hardened cement
to be destroyed (San et al., 2014). Therefore, GPC exhibits an
excellent high-temperature resistance. Through the test, it can be
known that GPC can withstand the high temperature of at least
800e1000  C(Nuaklong et al., 2020). Fig. 39 shows the change in
the compressive strength for OPCC, FA/GGBFS-based GPC, and FA/
OPC-based GPC exposed to temperatures ranging from 25 to
1100  C (Saavedra and Gutie rrez, 2017). The results indicate that at
the ambient temperature, the strength decreased in the following
order: FA/GGBFS-based GPC, OPCC, and FA/OPC-based GPC. For all
the specimens, the strength decreased with an increase in the
exposure temperature. However, the decline rate of the GPC was
Fig. 33. Carbonation changes for GPC and OPCC (Bernal et al., 2011). lower than that of the OPC, indicating that the GPC had an excellent
high-temperature resistance.
The bond performance of reinforced GPC embedded with plain
steel bars (GPC-1) and ribbed steel bars (GPC-2) were evaluated by
Zhang et al. (2018b). Fig. 40 presents the relationship between the
bond strength of the GPC specimens and the temperature. As is
shown, for all the samples, the bond strength changed slightly
when the temperature was below 300  C. However, the bond
strength of the samples was significantly reduced when the tem-
perature exceeded 300  C, possibly owing to the change in the
structure of the geopolymer caused by the high temperature, as
well as the thermal incompatibility of the geopolymer and the
aggregate. This trend is consistent with the compressive strength
and split tensile strength of GPC. The results indicated that 300  C
was the critical value for a large change in the bond strength of the
samples.
As is shown in Fig. 40, the bond strength of GPC with ribbed
steel bars at the ambient temperature was more than twice that of
GPC with plain steel bars, and the degradation rate of GPC-1 was
Fig. 34. Carbonation depth of GPC after 28 d of accelerated carbonation (Leemann higher than that of GPC-2. The bonding between the geopolymer
et al., 2015). and the plain steel bars mainly involved chemical adhesion and
friction. With an increase in the exposure temperature, the GPC was
fracture toughness of the GPC decreased. Additionally, the fractal severely deteriorated, which resulted in deformation and in-
dimension decreased when the slag/activator solution (S/A) ratio or compatibility between the geopolymer and the plain steel bars,
slag content increased, indicating that increasing the S/A ratio or reducing the friction (Zhang et al., 2016). Therefore, the bonding
slag content can improve the toughening of GPC. mechanism explains why the plain steel bars exhibited a lower
Shahrajabian and Behfamia (2018) examined the influence of NS pullout resistance than the ribbed steel bars. For the samples with
particles on the compressive strength of GGBFS-based GPC under ribbed steel bars, the mechanical occlusion between the geo-
freezeethaw cycling, as is shown in Fig. 38(a). The graph shows the polymer and the steel bars had a dominant effect on the bond
effect of the NS content on the compressive strength of the GPC. The strength. Therefore, the GPC-2 samples exhibited slower degrada-
NS significantly affected the compressive strength of the GPC under tion of the bond strength.
different freezeethaw conditions. Compared with the samples Fig. 41(a) and (b) show classical bond stresseslip curves for
without NS, the samples with NS exhibited a higher compressive specimens with plain steel bars and ribbed steel bars, respectively,
strength under the freezeethaw condition. Fig. 38(b) shows the at different temperatures. All the specimens failed owing to the
mass loss of GPC according to the NS content. The GPC samples pullout of the steel bars. The bond stress of GPC with ribbed steel
experienced mass loss during the freezeethaw cycles, because of bars was superior to that of GPC with plain steel bars. The bond
the damage to the edge of the specimens. The results indicated that stress of the GPC-2 samples decreased significantly when the
adding 3% NS to the GPC significantly reduced the mass loss of the temperature exceeded 300  C, and the bond stress of the GPC with
specimens. plain steel bars decreased more uniformly with the increasing
P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389 17

Fig. 35. Effect of MPCMs on the compressive strength (Pilehvar et al., 2019).

temperature (Zhang et al., 2018b).


Due to the high alkali concentration in the pores of FA/GGBFS-
based GPC, alkalis precipitate on the surface with the evaporation
of water in the process of condensation and hardening, frost on the
surface, and affect the durability of GPC (Maiti et al., 2020). In
particular, it affects the interface bonding properties of blocks.
Therefore, how to fix the alkali metal ions in the reaction products
to prevent the migration with water vapor is also a major problem.

5.5. Opportunities and challenges for FA/GGBFS-based GPC

Studies on the reaction mechanism and workability of FA/


GGBFS-based GPC have achieved tremendous progress. FA/
GGBFS-based GPC is ecofriendly with regard to production and
application, mainly because of the low CO2 emissions and the small
amount of water needed owing to the high FA content. The per-
formance of GPC is excellent without a superplasticizer (which is
expensive compared with concrete and increases the CO2 emis-
sions). Additionally, geopolymer technology provides opportunities
Fig. 36. Fracture toughness vs. fractal dimension (Cai et al., 2017). for recycling industrial waste, and this waste cannot be used in

Fig. 37. S/A ratio and slag content vs. fractal dimension for GPC (Cai et al., 2017).
18 P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389

Fig. 38. Compressive strength and mass loss of GPC exposed to freezeethaw cycles (Shahrajabian and Behfamia, 2018).

Fig. 39. Compressive strength of OPCC and GPC after exposure to high temperatures
rrez, 2017).
(Saavedra and Gutie

Fig. 40. Bond strength of GPC with plain and ribbed steel bars.
OPCC production. The foregoing factors satisfy the societal
requirement of sustainable development. Additionally, FA/GGBFS-
based GPC has broad application prospects because of its many developed for decades. Compared with traditional concrete, GPC
excellent characteristics, such as its shrinkage, corrosion resistance, has been studied less extensively, and the quantitative data on
and high-temperature resistance. its durability and mechanical properties are lacking.
Although the aforementioned factors have promoted the  Although GPC has excellent compressive strength, its tensile
development of FA/GGBFS-based GPC technology, GPC is seldom strength is weak, and brittle failure occurs easily. A toughening
used in large-scale engineering projects. There are obstacles for effect can be achieved by using a fiber-reinforced material to
achieving industrial application, which are summarized as follows: obtain the necessary mechanical and thermal properties for a
specific application.
 Published researched has confirmed the excellent properties of  The construction and manufacturing industries are conservative
FA/GGBFS-based GPC. However, geopolymers have different raw with regard to the adoption of new products. In developed
materials, and a large investment is required for the formulation countries, there are very specific standards for the performance
and certification of the raw materials. Although the changes of of the binder. Therefore, products such as GPC may not be
the raw materials are taken into account in the mix design, there completely acceptable, because GPC does not fully satisfy the
are difficulties in the process of implementation. regulatory standards, particularly with regard to the rheology
 There are various codes and standards for traditional concrete. and chemical composition. This can be a significant obstacle to
These relationships among strengths are based on extensive the adoption of GPC.
data regarding different engineering properties and have been
P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389 19

Fig. 41. Bond stresseslip curves for GPC reinforced with steel bars (Zhang et al., 2018b).

 Geopolymers can satisfy and exceed most of the existing per- alkali-activated material. Therefore, FA/GGBFS-based GPC has
formance standards in building applications, particularly when excellent freezeethaw resistance and high-temperature resis-
acid resistance and heat resistance are required. However, the tance. However, more investigations on the behavior at elevated
high shrinkage rate can only be improved by changing the temperature and the deterioration caused by environmental
curing conditions and using appropriate raw materials. factors of FA/GGBFS-based GPC are required.

The properties of FA/GGBFS-based GPC depend on the chemical


6. Conclusions composition of the binders, curing conditions, casting processes,
and environmental conditions. The review provides a foundation
The reaction mechanism of geopolymers, the workability, me- for understanding the effects of known parameters on the me-
chanical properties and durability of fresh and hardened FA/GGBFS- chanical properties and durability of fresh and hardened FA/GGBFS-
based GPC were reviewed. Based on the review results, the based GPC.
following conclusions can be drawn:

Declaration of competing interest


 Because of the main raw materials used in GPC are industrial
wastes, such as FA and GGBFS, the application of GPC can reduce
The authors declare that there is no conflict of interests in this
CO2 emissions, facilitate waste recycling and promote the sus-
paper.
tainable development of society. Therefore, the FA/GGBFS-based
GPC could be utilized potentially as a replacement for OPC.
However, this will only occur when there is an effective raw Acknowledgements
material supply chain.
 The reaction mechanism of the FA/GGBFS geopolymer material The authors would like to acknowledge the financial support
involves the breaking and recombining of the SieO and AleO received from State Key Laboratory of High Performance Civil En-
bonds under the action of an alkaline activator, which results gineering Materials (No. 2016CEM011), National Natural Science
in the production of a hardened geopolymer with excellent Foundation of China (No. 51678534, 51979251), Program for Inno-
mechanical properties and durability. vative Research Team (in Science and Technology) in University of
 The rheology of GPC differs from that of OPCC. Owing to the high Henan Province of China (No. 20IRTSTHN009), Changjiang River
viscosity of the Na2SiO3 solution, the slump of GPC decreases Scientific Research Institute (CRSRI) Open Research Program (No.
with an increase in the Na2SiO3 content. The SS/SH ratio, NaOH CKWV2018477/KY) and Open Projects Funds of Dike Safety and
concentration, and FA/GGBFS ratio affect the workability, initial Disaster Prevention Engineering Technology Research Center of
setting time, and final setting time of GPC. Additionally, the Chinese Ministry of Water Resources (No. 2018006).
workability of GPC with different mix proportions fluctuates
greatly. Therefore, the preparation technique of GPC with stable References
workability is the key problem in the promotion of GPC
applications. ACI Committee 318, 2011. Building Code Requirements for Structural Concrete and
Commentary. American Concrete Instute, Farmington Hills.
 The effects of different phases on the mechanical properties of Adak, D., Mandal, S., 2019. Strength and durability performance of fly ashebased
FA/GGBFS-based GPC were discussed. Comparisons indicated process-modified geopolymer concrete. J. Mater. Civ. Eng. 31, 04019174.
that FA/GGBFS-based GPC has better mechanical properties than AlBakri, A.M., Kamarudin, H., Bnhussain, M., Nizar, I.K., Rafiza, A., Zarina, Y., 2012.
The processing, characterization, and properties of fly ash based geopolymer
OPCC. SEM revealed that the structure of the geopolymer is concrete. Rev. Adv. Mater. Sci. 30, 90e97.
denser than that of OPC. XRD analysis indicated that GPC has a Alexander, M., Bertron, A., Belie, N., 2013. Performance of Cement-Based Materials
high crystallinity and that the ITZ of GPC is denser than that of in Aggressive Aqueous Environments. Springer, Netherlands.
Alonso, M., Gismera, S., Blanco, M., Lanzo n, M., Puertas, F., 2017. Alkali-activated
OPCC. However, the ITZ is the weakest part of OPCC. The FA/ mortars: workability and rheological behaviour. Construct. Build. Mater. 145,
GGBFS-based GPC poses desirable mechanical and structural 576e587.
properties that make it an ideal choice for the construction Alonso, S., Palomo, A., 2001. Calorimetric study of alkaline activation of calcium
hydroxideemetakaolin solid mixtures. Cement Concr. Res. 31, 25e30.
industry.
Ariffin, M., Bhutta, M., Hussin, M., Tahir, M.M., Aziah, N., 2013. Sulfuric acid resis-
 It can be indicated that FA/GGBFS-based GPC had a smaller tance of blended ash geopolymer concrete. Construct. Build. Mater. 43, 80e86.
strength loss and mass loss than OPCC after acid attack, sug- Aydın, S., Baradan, B., 2013. The effect of fiber properties on high performance
gesting that GPC has excellent acid resistance. However, the alkali-activated slag/silica fume mortars. Compos. B Eng. 45, 63e69.
iter, H., Baradan, B., 2007. Sulfuric acid resistance of high-
Aydın, S., Yazıcı, H., Yig
carbonation resistance of GPC is lower than that of OPCC. The volume fly ash concrete. Build. Environ. 42, 717e721.
structure of GPC is similar to that of zeolite because GPC is an Azzawi, M., Yu, T., Hadi, M.N., 2018. Factors affecting the bond strength between the
20 P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389

fly ash-based geopolymer concrete and steel reinforcement. Structure 262e272. Hadi, M.N., Farhan, N.A., Sheikh, M.N., 2017. Design of geopolymer concrete with
Bajpai, R., Choudhary, K., Srivastava, A., Sangwan, K.S., Singh, M., 2020. Environ- GGBFS at ambient curing condition using Taguchi method. Construct. Build.
mental impact assessment of fly ash and silica fume based geopolymer con- Mater. 140, 424e431.
crete. J. Clean. Prod. 254, 120147. Hanjitsuwan, S., Hunpratub, S., Thongbai, P., Maensiri, S., Sata, V., Chindaprasirt, P.,
Bakharev, T., Sanjayan, J.G., Cheng, Y.B., 2003. Resistance of alkali-activated slag 2014. Effects of NaOH concentrations on physical and electrical properties of
concrete to acid attack. Cement Concr. Res. 33, 1607e1611. high calcium fly ash geopolymer paste. Cement Concr. Compos. 45, 9e14.
Behfarnia, K., Rostami, M., 2017. An assessment on parameters affecting the Hansen, S.G., Lauridsen, J.T., Hoang, L.C., 2018. Experimental and statistical inves-
carbonation of alkali-activated slag concrete. J. Clean. Prod. 157, 1e9. tigation of the compressive strength anisotropy in structural concrete. Cement
Bernal, S.A., 2015. Effect of the activator dose on the compressive strength and Concr. Res. 107, 304e316.
accelerated carbonation resistance of alkali silicate-activated slag/metakaolin Hardjito, D., 2005. Studies of Fly Ash-Based Geopolymer Concrete. Curtin University.
blended materials. Construct. Build. Mater. 98, 217e226. Hardjito, D., Wallah, S.E., Sumajouw, D.M., Rangan, B.V., 2005. Fly ash-based geo-
Bernal, S.A., Gutie rrez, R.M., Pedraza, A.L., Provis, J.L., Rodriguez, E.D., Delvasto, S., polymer concrete. Aust. J. Struct. Eng. 6, 77e86.
2011. Effect of binder content on the performance of alkali-activated slag con- Ibrahim, M., Johari, M.A., Rahman, M.K., Maslehuddin, M., Mohamed, H.D., 2018.
cretes. Cement Concr. Res. 41, 1e8. Enhancing the engineering properties and microstructure of room temperature
Bhanja, S., Sengupta, B., 2005. Influence of silica fume on the tensile strength of cured alkali activated natural pozzolan based concrete utilizing nanosilica.
concrete. Cement Concr. Res. 35, 743e747. Construct. Build. Mater. 189, 352e365.
Cai, W., Cen, G., Wang, H., 2017. Fracture surface fractal characteristics of alkali-alag Juenger, M., Winnefeld, F., Provis, J.L., Ideker, J., 2011. Advances in alternative
concrete under freeze-thaw cycles. Ann. Mater. Sci. Eng. 3, 1e9. cementitious binders. Cement Concr. Res. 41, 1232e1243.
Castel, A., Foster, S.J., 2015. Bond strength between blended slag and Class F fly ash Junaid, M.T., Khennane, A., Kayali, O., Sadaoui, A., Picard, D., Fafard, M., 2014. Aspects
geopolymer concrete with steel reinforcement. Cement Concr. Res. 72, 48e53. of the deformational behaviour of alkali activated fly ash concrete at elevated
Cheah, C.B., Samsudin, M.H., Ramli, M., Part, W.K., Tan, L.E., 2017. The use of high temperatures. Cement Concr. Res. 60, 24e29.
calcium wood ash in the preparation of Ground Granulated Blast Furnace Slag Kastiukas, G., Ruan, S., Liang, S., Zhou, X., 2020. Development of precast geopolymer
and Pulverized Fly Ash geopolymers: a complete microstructural and me- concrete via oven and microwave radiation curing with an environmental
chanical characterization. J. Clean. Prod. 156, 114e123. assessment. J. Clean. Prod. 255, 120290.
Chen, X., Guo, Y., Ding, S., Zhang, H., Xia, F., Wang, J., Zhou, M.K., 2019. Utilization of Khater, H.M., 2014. Studying the effect of thermal and acid exposure on alkali
red mud in geopolymer-based pervious concrete with function of adsorption of activated slag geopolymer. Adv. Cement Res. 26, 1e9.
heavy metal ions. J. Clean. Prod. 207, 789e800. Krivenko, P., Drochytka, R., Gelevera, A., Kavalerova, E., 2014. Mechanism of pre-
Chi, M., Huang, R., 2013. Binding mechanism and properties of alkali-activated fly venting the alkalieaggregate reaction in alkali activated cement concretes.
ash/slag mortars. Construct. Build. Mater. 40, 291e298. Cement Concr. Compos. 45, 157e165.
Chindaprasirt, P., Silva, P., Sagoe, K., Hanjitsuwan, S., 2012. Effect of SiO2 and Al2O3 Kumar, S., Kumar, R., Mehrotra, S., 2010. Influence of granulated blast furnace slag
on the setting and hardening of high calcium fly ash-based geopolymer sys- on the reaction, structure and properties of fly ash based geopolymer. J. Mater.
tems. J. Mater. Sci. 47, 4876e4883. Sci. 45, 607e615.
Dahou, Z., Castel, A., Noushini, A., 2016. Prediction of the steel-concrete bond Kupwade-Patil, K., Allouche, E.N., 2012. Impact of alkali silica reaction on fly ash-
strength from the compressive strength of Portland cement and geopolymer based geopolymer concrete. J. Mater. Civ. Eng. 25, 131e139.
concretes. Construct. Build. Mater. 119, 329e342. L€
ammlein, T.D., Messina, F., Wyrzykowski, M., Terrasi, G.P., Lura, P., 2019. Low
Darwin, D., Dolan, C.W., Nilson, A.H., 2016. Design of Concrete Structures. McGraw- clinker high performance concretes and their potential in CFRP-prestressed
Hill Education. structural elements. Cement Concr. Compos. 100, 130e138.
Davidovits, J., Cordi, S., 1979. Synthesis of new high temperature geo-polymers for Laskar, S.M., Talukdar, S., 2017. Preparation and tests for workability, compressive
reinforced plastics/composites. SPE PACTEC 79, 151e154. and bond strength of ultra-fine slag based geopolymer as concrete repairing
Davidovits, J., 1989. Geopolymers and geopolymeric materials. J. Therm. Anal. Cal- agent. Construct. Build. Mater. 154, 176e190.
orim. 35, 429e441. Leemann, A., Nygaard, P., Kaufmann, J., Loser, R., 2015. Relation between carbon-
Davidovits, J., 1991. Geopolymers: inorganic polymeric new materials. J. Therm. ation resistance, mix design and exposure of mortar and concrete. Cement
Anal. Calorim. 37, 1633e1656. Concr. Compos. 62, 33e43.
Diaz, E.I., Allouche, E.N., Vaidya, S., 2011. Mechanical properties of fly-ash-based Lee, N., Lee, H.K., 2013. Setting and mechanical properties of alkali-activated fly ash/
geopolymer concrete. ACI Mater. J. 108, 300. slag concrete manufactured at room temperature. Construct. Build. Mater. 47,
Ding, Y., Bai, Y.L., 2018. Fracture properties and softening curves of steel fiber- 1201e1209.
reinforced slag-based geopolymer mortar and concrete. Materials 11, 1445. Lee, W., Van, J., 2007. Chemical interactions between siliceous aggregates and low-
Ding, Y., Dai, J.G., Shi, C.J., 2016. Mechanical properties of alkali-activated concrete: a Ca alkali-activated cements. Cement Concr. Res. 37, 844e855.
state-of-the-art review. Construct. Build. Mater. 127, 68e79. Le, H.T., Poh, L.H., Wang, S., Zhang, M.H., 2017. Critical parameters for the
Ding, Y., Shi, C.J., Li, N., 2018. Fracture properties of slag/fly ash-based geopolymer compressive strength of high-strength concrete. Cement Concr. Compos. 82,
concrete cured in ambient temperature. Construct. Build. Mater. 190, 787e795. 202e216.
Duxson, P., Lukey, G., Separovic, F., Van, J., 2005. Effect of alkali cations on aluminum Li, N., Shi, C.J., Zhang, Z.H., Zhu, D.J., Hwang, H.J., Zhu, Y.H., Sun, T.J., 2018. A mixture
incorporation in geopolymeric gels. Ind. Eng. Chem. Fundam. 44, 832e839. proportioning method for the development of performance-based alkali-acti-
Embong, R., Kusbiantoro, A., Shafiq, N., Nuruddin, M.F., 2016. Strength and micro- vated slag-based concrete. Cement Concr. Compos. 93, 163e174.
structural properties of fly ash based geopolymer concrete containing high- Li, Z., Zhang, S., Zuo, Y., Chen, W., Ye, G., 2019. Chemical deformation of metakaolin
calcium and water-absorptive aggregate. J. Clean. Prod. 112, 816e822. based geopolymer. Cement Concr. Res. 120, 108e118.
Fang, G., Ho, W.K., Tu, W., Zhang, M., 2018. Workability and mechanical properties of Liew, Y.M., Heah, C.Y., Kamarudin, H., 2016. Structure and properties of clay-based
alkali-activated fly ash-slag concrete cured at ambient temperature. Construct. geopolymer cements: a review. Prog. Mater. Sci. 83, 595e629.
Build. Mater. 172, 476e487. Lin, T., Jia, D., He, P., Wang, M., Liang, D., 2008. Effects of fiber length on mechanical
Farhan, N.A., Sheikh, M.N., Hadi, M.N., 2018. Experimental investigation on the ef- properties and fracture behavior of short carbon fiber reinforced geopolymer
fect of corrosion on the bond between reinforcing steel bars and fibre rein- matrix composites. Mater. Sci. Eng. 497, 181e185.
forced geopolymer concrete. Structure 251e261. Maiti, M., Sarkar, M., Maiti, S., Malik, M.A., Xu, S., 2020. Modification of geopolymer
Ferdous, W., Manalo, A., Khennane, A., Kayali, O., 2015. Geopolymer concrete-filled with size controlled TiO2 nanoparticle for enhanced durability and catalytic dye
pultruded composite beamseconcrete mix design and application. Cement degradation under UV light. J. Clean. Prod. 255, 120183.
Concr. Compos. 58, 1e13. Manjunath, R., Narasimhan, M.C., Umesha, K., 2019. Studies on high performance
Ferna ndez, A., Palomo, A., Vazquez, T., Vallepu, R., Terai, T., Ikeda, K., 2008. Alkaline alkali activated slag concrete mixes subjected to aggressive environments and
activation of blends of metakaolin and calcium aluminate. J. Am. Ceram. Soc. 91, sustained elevated temperatures. Construct. Build. Mater. 229, 116887.
1231e1236. Maranan, G., Manalo, A., Karunasena, K., Benmokrane, B., 2015. Bond stress-slip
Ferna ndez, A., Palomo, A., Sobrados, I., Sanz, J., 2006. The role played by the reactive behavior: case of GFRP bars in geopolymer concrete. J. Mater. Civ. Eng. 27,
alumina content in the alkaline activation of fly ashes. Microporous Meso- 04014116.
porous Mater. 91, 111e119. Mehta, A., Siddique, R., 2017a. Properties of low-calcium fly ash based geopolymer
Fernando, P.T., Joa ~o, C.G., Said, J., 2010. Durability and environmental performance concrete incorporating OPC as partial replacement of fly ash. Construct. Build.
of alkali-activated tungsten mine waste mud mortars. J. Mater. Civ. Eng. 22, Mater. 150, 792e807.
897e904. Mehta, A., Siddique, R., 2017b. Strength, permeability and micro-structural char-
Fernando, P.T., Said, J., 2011. Resistance to acid attack, abrasion and leaching acteristics of low-calcium fly ash based geopolymers. Construct. Build. Mater.
behavior of alkali-activated mine waste binders. Mater. Struct. 44, 487e498. 141, 325e334.
Glukhovsky, V., 2011. New cements for the 21st century: the pursuit of an alter- Meyer, C., 2009. The greening of the concrete industry. Cement Concr. Compos. 31,
native to Portland cement. Cement Concr. Res. 41, 750e763. 601e605.
Go€rhan, G., Kürklü, G., 2014. The influence of the NaOH solution on the properties of Moon, J., Bae, S., Celik, K., Yoon, S., Kim, K.H., Kim, K.S., Monteiro, P.J.M., 2014.
the fly ash-based geopolymer mortar cured at different temperatures. Compos. Characterization of natural pozzolan-based geopolymeric binders. Cement
B Eng. 58, 371e377. Concr. Compos. 53, 97e104.
Gu, L., Visintin, P., Bennett, T., 2018. Evaluation of accelerated degradation test Nath, P., Sarker, P.K., 2014. Effect of GGBFS on setting, workability and early strength
methods for cementitious composites subject to sulfuric acid attack; applica- properties of fly ash geopolymer concrete cured in ambient condition.
tion to conventional and alkali-activated concretes. Cement Concr. Compos. 87, Construct. Build. Mater. 66, 163e171.
187e204. Nath, P., Sarker, P.K., 2015. Use of OPC to improve setting and early strength
P. Zhang et al. / Journal of Cleaner Production 270 (2020) 122389 21

properties of low calcium fly ash geopolymer concrete cured at room temper- as a determining factor in the alkaline activation of aluminosilicates. Cement
ature. Cement Concr. Compos. 55, 205e214. Concr. Res. 42, 1242e1251.
Nath, P., Sarker, P.K., 2017. Flexural strength and elastic modulus of ambient-cured Saavedra, W.G., Gutie rrez, R.M., 2017. Performance of geopolymer concrete
blended low-calcium fly ash geopolymer concrete. Construct. Build. Mater. 130, composed of fly ash after exposure to elevated temperatures. Construct. Build.
22e31. Mater. 154, 229e235.
Noushini, A., Castel, A., Aldred, J., Rawal, A., 2020. Chloride diffusion resistance and San, R., Bernal, S.A., Gutie rrez, R.M., Deventer, J.S., Provis, J.L., 2014. Distinctive
chloride binding capacity of fly ash-based geopolymer concrete. Cement Concr. microstructural features of aged sodium silicate-activated slag concretes.
Compos. 105, 103290. Cement Concr. Res. 65, 41e51.
Nuaklong, P., Sata, V., Chindaprasirt, P., 2016. Influence of recycled aggregate on fly Shahrajabian, F., Behfarnia, K., 2018. The effects of nano particles on freeze and thaw
ash geopolymer concrete properties. J. Clean. Prod. 112, 2300e2307. resistance of alkali-activated slag concrete. Construct. Build. Mater. 176,
Nuaklong, P., Jongvivatsakul, P., Pothisiri, T., Sata, V., Chindaprasirt, P., 2020. Influ- 172e178.
ence of rice husk ash on mechanical properties and fire resistance of recycled Shi, X., Collins, F.G., Zhao, X.L., Wang, Q., 2012. Mechanical properties and micro-
aggregate high-calcium fly ash geopolymer concrete. J. Clean. Prod. 252, 119797. structure analysis of fly ash geopolymeric recycled concrete. J. Hazard Mater.
Okoye, F.N., Prakash, S., Singh, N.B., 2017. Durability of fly ash based geopolymer 237, 20e29.
concrete in the presence of silica fume. J. Clean. Prod. 149, 1062e1067. Singhal, D., Junaid, M.T., Jindal, B.B., Mehta, A., 2018. Mechanical and microstruc-
Pan, Z., Sanjayan, J.G., Rangan, B.V., 2011. Fracture properties of geopolymer paste tural properties of fly ash based geopolymer concrete incorporating alccofine at
and concrete. Mag. Concr. Res. 63, 763e771. ambient curing. Construct. Build. Mater. 180, 298e307.
Pasupathy, K., Berndt, M., Sanjayan, J., Rajeev, P., Cheema, D.S., 2017. Durability of Sofi, M., Van, J., Mendis, P., Lukey, G., 2007. Engineering properties of inorganic
low-calcium fly ash based geopolymer concrete culvert in a saline environment. polymer concretes (IPCs). Cement Concr. Res. 37, 251e257.
Cement Concr. Res. 100, 297e310. Somna, K., Jaturapitakkul, C., Kajitvichyanukul, P., Chindaprasirt, P., 2011. NaOH-
Peng, J.X., Huang, L., Zhao, Y.B., Chen, P., Zeng, L., 2012. Modeling of carbon dioxide activated ground fly ash geopolymer cured at ambient temperature. Fuel 90,
measurement on cement plants. Adv. Mater. Res. 610e613, 2120e2128. 2118e2124.
Phoongernkham, T., Chindaprasirt, P., Sata, V., Hanjitsuwan, S., Hatanaka, S., 2014. Song, P., Hwang, S., 2004. Mechanical properties of high-strength steel fiber-
The effect of adding nano-SiO2 and nano-Al2O3 on properties of high calcium fly reinforced concrete. Construct. Build. Mater. 18, 669e673.
ash geopolymer cured at ambient temperature. Mater. Des. 58e65. Tuinukuafe, A., Kaub, T., Weiss, C.A., Allison, P.G., Thompson, G.B., Amirkhanian, A.,
Pilehvar, S., Szczotok, A.M., Rodríguez, J.F., Valentini, L., Lanzo n, M., Pamies, R., 2019. Atom probe tomography of an alkali activated fly ash concrete. Cement
Kjoniksen, A.L., 2019. Effect of freeze-thaw cycles on the mechanical behavior of Concr. Res. 121, 37e41.
geopolymer concrete and Portland cement concrete containing micro- Venu, M., Rao, T.G., 2017. Tie-confinement aspects of fly ash-GGBS based geo-
encapsulated phase change materials. Construct. Build. Mater. 200, 94e103. polymer concrete short columns. Construct. Build. Mater. 151, 28e35.
Prasittisopin, L., Sereewatthanawut, I., 2018. Effects of seeding nucleation agent on Wang, Y., Hu, S., He, Z., 2019. Mechanical and fracture properties of fly ash geo-
geopolymerization process of fly-ash geopolymer. Front. Struct. Civ. Eng. 12, polymer concrete addictive with calcium aluminate cement. Materials 12, 2982.
16e25. Won, J.P., Kim, J.H., Park, C.G., Kang, J.W., Kim, H.Y., 2009. Shrinkage cracking of
Provis, J.L., Deventer, J.S., 2009. Geopolymers: Structures, Processing, Properties and styrene butadiene polymeric emulsion-modified concrete using rapid-hard-
Industrial Applications. Springer, Netherlands. ening cement. J. Appl. Polym. Sci. 112, 2229e2234.
Puertas, F., Gonza lez, B., Gonza lez, I., Alonso, M., Torres, M., Rojo, G., Martinez- Wongsa, A., Zaetang, Y., Sata, V., Chindaprasirt, P., 2016. Properties of lightweight fly
Abella, F., 2018. Alkali-activated slag concrete: fresh and hardened behaviour. ash geopolymer concrete containing bottom ash as aggregates. Construct. Build.
Cement Concr. Compos. 85, 22e31. Mater. 111, 637e643.
Puertas, F., Palacios, M., Manzano, H., Dolado, J., Rico, A., Rodríguez, J., 2011. A model Xie, N., Dang, Y., Shi, X., 2019. New insights into how MgCl2 deteriorates Portland
for the CASH gel formed in alkali-activated slag cements. J. Eur. Ceram. Soc. 31, cement concrete. Cement Concr. Res. 120, 244e255.
2043e2056. Yang, K., Yang, C., Magee, B., Nanukuttan, S., Ye, J., 2016. Establishment of a pre-
Puertas, F., Varga, C., Alonso, M.M., 2014. Rheology of alkali-activated slag pastes. conditioning regime for air permeability and sorptivity of alkali-activated slag
Effect of the nature and concentration of the activating solution. Cement Concr. concrete. Cement Concr. Compos. 73, 19e28.
Compos. 53, 279e288. Yip, C.K., Deventer, J.S., 2003. Microanalysis of calcium silicate hydrate gel formed
Puligilla, S., Mondal, P., 2013. Role of slag in microstructural development and within a geopolymeric binder. J. Mater. Sci. 38, 3851e3860.
hardening of fly ash-slag geopolymer. Cement Concr. Res. 43, 70e80. Yip, C.K., Lukey, G., Deventer, J.S., 2005. The coexistence of geopolymeric gel and
Qian, L., Wang, Y., Alrefaei, Y., Dai, J., 2020. Experimental study on full-volume fly calcium silicate hydrate at the early stage of alkaline activation. Cement Concr.
ash geopolymer mortars: sintered fly ash versus sand as fine aggregates. Res. 35, 1688e1697.
J. Clean. Prod., 121445 Zhang, H.Y., Kodur, V., Wu, B., Cao, L., Wang, F., 2016. Thermal behavior and me-
Ranjbar, N., Talebian, S., Mehrali, M., Kuenzel, C., Metselaar, H.S.C., Jumaat, M.Z., chanical properties of geopolymer mortar after exposure to elevated temper-
2016. Mechanisms of interfacial bond in steel and polypropylene fiber rein- atures. Construct. Build. Mater. 109, 17e24.
forced geopolymer composites. Compos. Sci. Technol. 122, 73e81. Zhang, P., Zheng, Y.X., Wang, K.J., Zhang, J.P., 2018a. A review on properties of fresh
Rao, G., Tippabhotla, D., Alfrite, P., Mallikarjuna, G., Andal, M., 2015. Fracture pa- and hardened geopolymer mortar. Compos. B Eng. 152, 79e95.
rameters of fly ash and GGBS based geopolymer concrete. Appl. Mech. Mater. Zhang, H.Y., Kodur, V., Wu, B., Yan, J., Yuan, Z.S., 2018b. Effect of temperature on
764e765, 1090e1094. bond characteristics of geopolymer concrete. Construct. Build. Mater. 163,
Rattanasak, U., Chindaprasirt, P., 2009. Influence of NaOH solution on the synthesis 277e285.
of fly ash geopolymer. Miner. Eng. 22, 1073e1078. Zhang, Y.J., Li, H.H., Wang, Y.C., Xu, D.L., 2011. Geopolymer microstructure and hy-
Razak, A., Nooriza, S., Nuruddin, M., 2014. The effects of alkali silica reaction (ASR) dration mechanism of alkali-activated fly ash-based geopolymer. Adv. Mater.
towards fly ash based geopolymer concrete. Appl. Mech. Mater. 699, 271e276. Res. 374, 1481e1484.
Rees, C.A., Provis, J.L., Lukey, G.C., Deventer, J.S., 2008. The mechanism of geo- Zhang, Z., Provis, J.L., Reid, A., Wang, H., 2014. Fly ash-based geopolymers: the
polymer gel formation investigated through seeded nucleation. Colloids Surf., A relationship between composition, pore structure and efflorescence. Cement
318, 97e105. Concr. Res. 64, 30e41.
Ronad, A., Karikatti, V., Dyavanal, S., 2016. A study on mechanical properties of Zhuang, X.Y., Chen, L., Komarneni, S., Zhou, C.H., Tong, D.S., Yang, H.M., Yu, W.H.,
geopolymer concrete reinforced with basalt fiber. Int. J. Res. Eng. Technol. 5, Wang, H., 2016. Fly ash-based geopolymer: clean production, properties and
474e478. applications. J. Clean. Prod. 125, 253e267.
Ruiz, C., Skibsted, J., Ferna ndez, A., Palomo, A., 2012. Alkaline solution/binder ratio

You might also like