Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Construction and Building Materials 293 (2021) 123527

Contents lists available at ScienceDirect

Construction and Building Materials


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

Effects of initial SiO2/Al2O3 molar ratio and slag on fly ash-based ambient
cured geopolymer properties
Ayoub Dehghani a, Farhad Aslani a,b,⇑, Neda Ghaebi Panah c
a
Materials and Structures Innovation Group, School of Engineering, University of Western Australia, WA 6009, Australia
b
School of Engineering, Edith Cowan University, WA 6027, Australia
c
School of Engineering, University of Western Australia, 35 Stirling Highway, Perth, WA 6009, Australia

h i g h l i g h t s

 The effect of the initial molar ratio of SiO2/Al2O3 is discussed.


 The effect of the fly ash replacement with GGBFS is discussed.
 The maximum compressive strength was achieved for a SiO2/Al2O3 ratio of 3.37.
 Amorphous fractions of all geopolymers were between 81% to 84%.
 BET surface area decreased by increasing SiO2/Al2O3 ratio or GGBFS content.

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

Article history: Geopolymers cured at ambient temperature are of high interest since they provide promising environ-
Received 20 February 2021 mentally friendly alternative to cement. This paper discusses the effect of the initial molar ratio of
Received in revised form 29 April 2021 SiO2/Al2O3 and fly ash replacement with ground granulated blast furnace slag (GGBFS) on the properties
Accepted 29 April 2021
of fly ash-based geopolymers. Geopolymer mixtures developed in this study are user-friendly (SiO2/Na2O
Available online 11 May 2021
larger than 1.4) and suitable for producing self-compacting geopolymer concrete at ambient temperature.
Results showed that the increase in SiO2/Al2O3 ratio accelerated the setting of fly ash-slag geopolymer
Keywords:
systems, which was quite different from that reported for metakaolin-based geopolymers. An increas-
Geopolymer
Molar ratio
ing–decreasing trend in compressive strength was observed by increasing the SiO2/Al2O3 ratio. The max-
Slag imum compressive strength achieved when the SiO2/Al2O3 ratio was 3.37 (Si/Al = 1.68). The higher
Fly ash compressive strength at this ratio could be related to the alumina-silica bondings in the amorphous area,
Ambient cured which comprised more than 81%-84% of geopolymer microstructure. Furthermore, the study revealed
Microstructure that the decrease in fly ash replacement with GGBFS prolonged the setting of geopolymer mixtures
and reduced their compressive strength significantly.
Ó 2021 Elsevier Ltd. All rights reserved.

1. Introduction One such progress is the development of geopolymer binders.


These binders are a promising alternative to OPC and can eliminate
Ordinary Portland cement (OPC) manufacturing is responsible the need for OPC in concrete production [8]. Geopolymers are inor-
for the emission of 5–7% of the global carbon dioxide emission ganic polymers with a broad application, including mortars and
[1–3], which is the main component in greenhouse gases and concrete, art and decorating materials, ceramic tiles, thermal insu-
accountable for 65% of global warming [4]. Moreover, the produc- lating materials, fire-resistant materials, and high-tech composites
tion of OPC is associated with depleting natural resources, such as [9]. Several investigations have been carried out to explore the syn-
raw limestone and fossil fuel [5]. 1.5 tonnes of raw material is thesis method and properties of geopolymers [10–13].
needed for the manufacturing of one tonne of OPC [4,6,7]. Hence, Geopolymers are synthesised by activating aluminosilicate
significant developments have been made in the past decades to materials such as calcined metakaolin [14], fly ash, slag, silica
partially or entirely replace OPC in concrete with other products. fume, and rice husk with alkaline solutions [15–17]. The mechan-
ical and microstructure properties of geopolymers are governed by
⇑ Corresponding author at: Materials and Structures Innovation Group, School of the chemical composition of source aluminosilicate materials and
Engineering, University of Western Australia, WA 6009, Australia alkaline solutions [11,18,19]. The alkaline solution can be prepared
E-mail address: farhad.aslani@uwa.edu.au (F. Aslani). using sodium hydroxide (NaOH), potassium hydroxide (KOH),

https://doi.org/10.1016/j.conbuildmat.2021.123527
0950-0618/Ó 2021 Elsevier Ltd. All rights reserved.
A. Dehghani, F. Aslani and N. Ghaebi Panah Construction and Building Materials 293 (2021) 123527

sodium silicate (Na2SiO3), and potassium silicate (K2SiO3) [13]. sation process comparing to metakaolin-based geopolymers [34].
Although NaOH has a lower level of alkalinity comparing to KOH, Hence, the effect of initial molar ratios on different properties of
it showed a better performance in liberating silicate and aluminate geopolymer pastes developed using by products (i.e., fly ash and
monomers during geopolymerisation process [11]. This process slag) have been investigated in the present study.
can be simplified in three phases, including dissolution- On the other hand, when it comes to GC, other variables such as
hydrolysis of Si-tetrahedra and Al-tetrahedra of raw aluminosili- the concentration of NaOH solution, sodium silicate to sodium
cate materials in the alkaline solution, the polycondensation of hydroxide ratio (SS/SH), the alkaline solution to binder ratio (A/
Si-tetrahedra and Al-tetrahedra into oligomers, followed by the b), and water to binder ratio (w/b) are commonly considered by
formation of amorphous to semi-crystalline 3D silico-aluminate researchers to study the mechanical properties of mixtures
structures called geopolymer [8,9,20]. Poly(sialate) with the fol- [8,32,35–42]. A wide range of these ratios was considered in previ-
lowing empirical formula has been suggested for the chemical des- ous studies to develop GC with different precursors; 2–18 for the
ignation of geopolymers [21]. concentration of NaOH solution, 0.4–4.0 for SS/SH, 0.30–0.94 for
 A/b, and 0.17–0.53 for w/b. The recommended values for these
M n ðSiO2 Þz  AlO2 n  wH2 O ratios to develop GC with better mechanical properties (i.e., mainly
compressive strength) were different in these studies, even among
where M is an alkali cation such as Na+, K+, and Ca+, which balances those used almost similar materials and curing conditions. For
the negative charge of Al3+ IV-fold coordination, n stands for the instance, Mustafa et al. [37] reported the highest compressive
degree of polycondensation, and z is the Si/Al molar ratio. strength for fly ash based GC cured at 70 °C with a 12 M NaOH con-
The chemistry of geopolymers is mainly affected by the type, centration and an SS/SH = 2.5. In comparison, fly ash-based
chemical composition, and reactivity of raw aluminosilicate mate- geopolymer mixes cured at 85 °C by Görhan and Kürklü [41] exhib-
rials and the chemical characteristics of alkaline solutions. Hence, ited higher compressive strength at SS/SH = 2.0 and 6 M NaOH.
the design of geopolymer mixes with predictable properties is far Another example can be found in a review study carried out by
from straightforward [5,22]. Various parameters such as the reac- Ng et al. [12] where a significant discrepancy observed in the com-
tive amount of SiO2 and Al2O3 in starting material, Na2O/SiO2 and pressive strength of geopolymer mixes with similar precursors.
H2O/Na2O ratio, synthesis condition, curing time, curing tempera- Therefore, it is more reasonable to examine the mechanical proper-
ture, and pH affect the properties of geopolymers [9,10]. However, ties of GC based on the initial molar ratios (i.e., SiO2/Al2O3, Na2O/
the molar ratio of SiO2/Al2O3 has a fundamental effect on setting Al2O3, H2O/Na2O) provided by precursor materials instead of the
time and strength development of geopolymer systems [23]. More factors mentioned above (i.e., the concentration of NaOH solution,
specifically, the presence of a low Si/Al ratio (=1) results mainly in SS/SH, A/b, etc.).Hence, this study aimed to investigate the effect of
the formation of poly(sialate) polymer structures. While increasing initial molar ratios on the setting time, mechanical properties, and
this ratio favours the formation of geopolymeric compounds of the microstructure of geopolymer paste suitable for producing self-
poly (sialate-siloxo) (Si/Al = 2), poly(sialate- disiloxo) (Si/Al = 3), compacting fly ash based GC at ambient temperature. Two vari-
and poly(sialate- multisiloxo) (Si/Al > 3) type [10,23–25]. ables considered were SiO2/Al2O3 ratio while Na2O/Al2O3 and
Several researchers reported the effect of initial molar ratio of H2O/Na2O were fixed and the amount of ground granulated blast
SiO2/Al2O3 on the mechanical properties and microstructure of furnace slag (GGBFS). Furthermore, the SiO2/Na2O ratio of alkaline
metakaolin-based geopolymers cured at different temperatures solutions was larger than 1.4 for almost all mixes to ensure user-
(40 °C to 90 °C) for various time (20 h to 7 months) [9,24,26–30]. friendly conditions suitable for mass production [43,44].
Most of these studies reported higher compressive strength for
geopolymers synthesised with a SiO2/Al2O3 ratio between 3.0 and
5.0 with an Na2O/Al2O3 ratio between 0.5 and 1.3. According to 2. Experimental program
results of these studies, the increase in the SiO2/Al2O3 ratio beyond
this range resulted in low strength products. For instance, Duxson 2.1. Materials
et al. [24] observed 400% increase in the compressive strength of
geopolymers by increasing the SiO2/Al2O3 ratio of the mix from Low-calcium fly ash class F as per AS/NZS 3582.1 [45] was used
2.3 to 3.8 before decreasing again at the largest studied SiO2/ as the primary aluminosilicate precursor with a fineness of 88%
Al2O3 ratio of 4.3. In comparison, Davidovits suggested SiO2/ passing a 45 mm sieve. The scanning electron microscopy (SEM)
Al2O3 ratios in the range of 3.5–4.5 (with a preferable value of of fly ash revealed its spherical and smooth-surface particles
4.0) and a Na2O/Al2O3 ratio between 0.8 and 1.2 for developing (Fig. 1a), which could increase the flowability of GC [36]. GGBFS
geopolymers [31]. It is noteworthy that the H2O/Na2O ratio of was used to facilitate ambient curing conditions and improve the
the mixture is also important to synthesise geopolymers success- mechanical properties of GC [8]. GGBFS was complied with AS
fully. Although this ratio can be changed in the range of 10–25 3582.2 [46] and characterised by irregular and angular particles,
[10], amorphous geopolymeric phase and higher compressive as shown in Fig. 1b. The specific gravity of FA and GGBFS were
strength were observed for H2O/Na2O ratio in the range of 10–15 2.29 and 3.10, respectively. The volume median diameter (D50) of
[9,23,27,29]. H2O/Na2O ratio was also suggested to be kept within FA and GGBFS obtained through the laser diffraction technique
the range of 10–14 to achieve workable geopolymer concrete (GC) were 7.8 mm and 11.2 mm, respectively. Table 1 lists the chemical
[32]. compositions and loss on ignition (LOI) of fly ash and GGBFS deter-
The initial molar ratios mentioned above may change if precur- mined by X-ray Fluorescence spectrometer (XRF). The filler of
sors other than metakaolin are used to provide SiO2 and Al2O3 for geopolymer paste consisted of fine silica sand with a maximum
geopolymerisation [23]. Furthermore, these ratios work with aggregate size of 600 mm to achieve a proper packing density.
highly reactive precursors such as metakaolin and nano-silica Fig. 2 shows the particle size distribution for FA and GGBFS
and nano-alumina [9,10,33]. Such materials are characterised by obtained by Malvern Mastersizer 2000 laser particle size analyzer
very high dissolution capacity in alkaline solution, resulting in final and for fine snad achieved through sieve analysis.
systems with similar molar ratios as initial precursors. However, The alkaline solutions used in this study were mixtures of liquid
suggested initial molar ratios may change when the use of by- N-grade sodium silicate and 14 M (560 g/L) sodium hydroxide,
products, in particular, fly ash, is considered [10]. Also, fly ash is which prepared by dissolving 99% pure sodium hydroxide pellets
highly heterogeneous, having a significant effect on geopolymeri- in water. The chemical composition of the sodium silicate was
2
A. Dehghani, F. Aslani and N. Ghaebi Panah Construction and Building Materials 293 (2021) 123527

Fig. 1. SEM image of (a) fly ash and (b) GGBFS.

Table 1
Chemical compositions of fly ash and GGBFS.

SiO2 Al2O3 Fe2O3 CaO MgO Na2O SO3 K2O P 2 O5 TiO2 Mn2O3 LOI Specific Fineness (passing
(%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) gravity 45 lm)
FA 50.4 31.5 10.4 3.3 1.1 0.3 0.2 0.5 0.5 – 0.2 1.1 2.29 88%
GGBFS 32.9 13.8 0.5 42.1 5.8 0.2 3.3 0.3 0.03 0.6 0.3 0.2 3.10 98%

influence of change in these parameters on final products. Note


that w/b ratio was calculated based on the total water of the
geopolymer system, including water from 14 M NaOH, sodium sil-
icate, and extra added water. It was also tried to develop mixtures
with a SiO2/Al2O3 ratio of 4.0 and 3.0 in group one. However, these
pastes were not practical to be implemented in geopolymer con-
crete due to their poor fresh properties (i.e., very high fluidity for
the former and very poor workability for the latter).
First, fly ash and GGBFS was mixed for 2 min by using a labora-
tory Hobart mixer. Then, the prepared alkaline solution was added
and mixed for 3 min, followed by adding water and mixing for
another 2 min. The mixing process was continued by adding fine
sand at the last stage and stirring for 3 min. The slump and setting
time tests were always started 2 min and 12 min after the mixing
process, respectively. Mixtures were cast in cubic moulds with an
edge length of 50 mm to prepare compressive samples. The speci-
mens were sealed with plastic to prevent water loss for 24 h, fol-
lowed by demoulding and placing them in a moist chamber at
23 ± 1 °C and 92 ± 2% RH until the day of the test.
Fig. 2. Particle-size distribution of fly ash, GGBFS, and fine sand.

2.3. Testing procedures


28.7% SiO2, 8.9% Na2O, and 62.4% H2O with silica molecular modu-
lus Ms = 3.2 and a density of 1.39 g/mL. First sodium hydroxide was The flowability of mixtures was examined by the miniature
added to water and mixed for 5 min, followed by about 15 min rest slump (mini-slump) test 10 min after adding the alkaline solution
time for the chemical reaction to take place. Then, the sodium sil- to the mix. The slump cone used had a height of 50 mm with a top
icate was added and mixed for another 5 min and left for 24 h and bottom diameter of 70 mm and 100 mm (d0 ) respectively. The
before mixing with the precursor [5]. slump cone positioned on a flat and levelled PVC plate was filled
with the fresh geopolymer paste and lifted upward, followed by
measuring the spread diameter in two perpendicular directions
2.2. Mixtures proportions, specimens preparation, and curing after 2 min. The average of the two measurements (d) was used
in the following formula to calculate the corresponding relative
The mixtures were divided into two main groups, as sum- slump (Cp ).
marised in Table 2. The first group was designed to study the effect
of SiO2/Al2O3 ratio with a constant Na2O/Al2O3 and H2O/Na2O ratio.  2
d
In comparison, the second group aimed to determine the influence Cp ¼ 1 ð1Þ
d0
of fly ash replacement with GGBFS regardless of the change in the
molar ratios of mixtures. The mixes categorized as group 2 had a The cubic specimens were tested under compression using a
constant w/b = 0.35, SS/SH = 2.5, and A/b = 0.40 to eliminate the Baldwin machine under a loading rate of 0.5 mm/min.
3
A. Dehghani, F. Aslani and N. Ghaebi Panah Construction and Building Materials 293 (2021) 123527

Table 2
Weight proportions of geopolymer mixtures with respect to the binder (i.e., FA + GGBFS) and corresponding molar ratios.

Group Sample Fly GGBFS Sodium NaOH solution Water* sand SiO2/Al2O3 (molar Na2O/Al2O3 (molar H2O/Na2O (molar
ash silicate 14 M ratio) ratio) ratio)
1 Si375 0.820 0.180 0.531 0.038 0.207 0.934 3.75 0.35 13.1
Si350 0.834 0.166 0.396 0.079 0.178 0.863 3.50 0.35 13.1
Si337 0.778 0.222 0.285 0.114 0.156 0.805 3.37 0.35 13.1
Si325 0.851 0.149 0.262 0.125 0.152 0.772 3.25 0.35 13.1
2 GS22 0.778 0.222 0.285 0.114 0.156 0.805 3.37 0.35 13.1
GS15 0.845 0.155 0.285 0.114 0.156 0.805 3.30 0.34 13.1
GS8 0.922 0.078 0.285 0.114 0.156 0.805 3.22 0.33 13.1
GS4 0.961 0.039 0.285 0.114 0.156 0.805 3.19 0.32 13.1

* Extra added water.

The initial and final time of setting of geopolymer mixes were that the increase of SiO2/Al2O3 ratio up to 3.75 accelerated settings
measured using Vicat needle apparatus as per ASTM C191 [47]. of fly ash-slag geopolymer systems. This is quite different from that
The initial contact of the precursors and the alkaline solution of conventional geopolymers (i.e., metakaolin-based geopolymers).
was considered as zero time. The Vicat initial time of setting was In conventional geopolymers, the increase in SiO2/Al2O3 ratio pro-
calculated as the time corresponding to a needle penetration of longed setting [23]. The results presented in Figs. 3a and 4a agree
25 mm. The elapsed time until the needle could leave a circular well with those reported by Chindaprasirt et al. [48] for high cal-
impression on the mixture was considered the final setting. The cium fly ash-based geopolymer systems cured at 60 °C. They
Vicat test was performed in the indoor environment with temper- showed that the increase in the SiO2 content of geopolymer with
ature and humidity of 23 ± 2 °C and 54 ± 2%. a fixed content of Al2O3 accelerated the setting of geopolymers.
The compression test was carried out at room temperature and For mixtures in group 2, Figs. 3b and 4b reveal that the increase
a constant quasi-static test speed of 0.5 mm/min using a Baldwin in fly ash replacement with GGBFS accelerated the setting of
universal testing machine. The compressive strength was tested geopolymer mixtures. This observation agrees with the published
at 1, 3, 7, 14, 28, 56 and 90 days age of geopolymer mixes. The findings that the existence of soluble calcium accelerates the set-
spherically-seated platen that existed between the top-loading ting of geopolymers [18,49–51]. It could be due to the precipitation
plate and specimens ensured a uniform stress distribution on spec- of calcium silicate hydrates providing nucleation sites that trigger
imens. Small broken pieces of compressive samples were collected quick geopolymerisation [18]. Aside from the evident effect of
for microstructure analyses using a Vega3 Tescan scanning elec- GGBFS replacement on setting times, the geopolymer mixtures in
tron microscope (SEM) at the accelerating voltage of 26 kV. group 2 revealed that setting time increases by decreasing SiO2/
Al2O3 ratio (see Table 2). This result from group 2 aligns with that
of group 1 regarding setting time. Note that for the mixes in group
3. Results and discussion
2, SiO2/Al2O3 ratio decreased by increasing GGBFS replacement
while Na2O/Al2O3 changed slightly. Hence, it can be concluded that
3.1. Setting time
SiO2/Al2O3 ratio governed the setting of geopolymers in both
groups. This conclusion is consistent with the findings of Nath
Fig. 3 shows the evolution of depth penetration of Vicat needle
and Sakar [52]. They reported an accelerated setting for geopoly-
for the geopolymer mixes at room temperature. Fig. 4 shows the
mer mixes synthesized using higher SS/SH. Indeed, increasing SS/
effect of SiO2/Al2O3 ratio and GGBFS replacement on the initial
SH caused a mixture with a greater SiO2/Al2O3 molar ratio.
and final setting time of geopolymers mixes. Figs. 3a and 4a reveal

Fig. 3. Evolution of depth penetration: (a) the effect of SiO2/Al2O3 ratio and (b) the effect of GGBFS replacement.

4
A. Dehghani, F. Aslani and N. Ghaebi Panah Construction and Building Materials 293 (2021) 123527

Fig. 4. The effect of (a) SiO2/Al2O3 ratio and (b) slag on Vicat initial and final time of setting.

3.2. Flowability also worth mentioning that the effect of GGBFS on the geopolymer
mixtures was opposite to its impact on cementitious mixes [55].
The effect of GGBFS replacement and SiO2/Al2O3 ratio on the Significant reduction in flowability was also observed by increasing
flowability of mixtures is shown in Fig. 5a and b, respectively. SiO2/Al2O3 ratio (see Figs. 5b and 6), which is consistent with the
Fig. 5a indicates that when GGBFS inclusion was increased, results observed for setting times. The mixtures with a higher
geopolymer mixtures showed lower flowability. The flow diameter SiO2/Al2O3 ratio were more reactive and generated geopolymer
decreased from 372 mm to 342 mm when the GGBFS content was networks more quickly. This could increase the viscosity of mix-
increased from 0.039 to 0.222 (see Fig. 6). Note that except for tures and reduce their flowability.
GGBFS to fly ash ratio, all parameters, including w/b and A/b, were
fixed for the mixtures shown in this figure. The reduction in flowa- 3.3. Compressive strength development
bility could be attributed to two reasons, particle shape and the
reactivity of the mixture. Increasing the replacement of fly ash by The compressive strength development results of geopolymers
GGBFS content increases GGBFS angular particles and decrease at different ages have been plotted in Fig. 7a and b for mixtures
spherical fly ash particles. The presence of a higher number of in group 1 and 2, respectively. The initial SiO2/Al2O3 ratio and
angular particles can reduce the flowability of the mixture. On GGBFS content had a significant effect on the compressive strength
the other hand, the increase in GGBFS content of the mixture leads of ambient cured mixtures. The compressive strength of all
to a more reactive mix, resulting in a faster setting and lower work- geopolymers increased as a function of the mixture age. The plot-
ability [53]. Note that the reaction between precursors and the ted graphs revealed that the compressive strength development of
alkaline solution starts as soon as they are mixed [54]. Since the all mixtures increased rapidly until 28 days, followed by a further
flow diameter was measured 10 min after adding alkaline solution, increase in compressive strength at slower rates until 90 days of
any change in reactivity rate can change the flow diameter. It is age, which is in agreement with published findings [9,39].

Fig. 5. The effect of (a) slag content ratio and (b) SiO2/Al2O3 ratio on the flowability of geopolymer mixtures.

5
A. Dehghani, F. Aslani and N. Ghaebi Panah Construction and Building Materials 293 (2021) 123527

Fig. 6. The flow of three selected geopolymer pastes 2 min after removal of mini-slump cone.

Fig. 7. The effect of (a) SiO2/Al2O3 ratio and (b) slag content on the compressive strength of geopolymers.

Fig. 8. X-ray diffractograms of geopolymer at 28 days.

6
A. Dehghani, F. Aslani and N. Ghaebi Panah Construction and Building Materials 293 (2021) 123527

Fig. 7a shows that the increase in SiO2/Al2O3 ratio from 3.25 to fly ash [56,57]. Various parameters such as temperature, alkalinity,
3.37 caused a significant improvement in compressive strength dissolved species, crystallinity, and structural defects have been
before decreasing again at a higher SiO2/Al2O3 ratio. The observed reported to affect the dissolution rate of aluminosilicate in fly
increasing–decreasing tendency of compressive strength by ash, slag, metakaolin, etc. [57,58]. Fig. 1a also suggests the SiO2/
increasing the SiO2/Al2O3 ratio was similar to that reported for Al2O3 ratio governs the growth of compressive strength at early
metakaolin-based geopolymers [24,26]. The maximum compres- ages. For instance, the compressive strength of Si337 was
sive strength at different geopolymer ages was achieved for a 11.7 MPa and 16.5 MPa at 1 day and 3 days ages, respectively,
SiO2/Al2O3 ratio of 3.37 (Si/Al = 1.68). This ratio was slightly lower while these figures were 2.5 MPa and 5.6 MPa for Si375.
than the optimum value (i.e., Si/Al = 1.9) reported by Duxson et al. The replacement of fly ash by GGBFS increased the compressive
[24] for metakaolin-based geopolymers cured at 40 °C, which could strength of geopolymers, as shown in Fig. 7. This increase was
be due to the higher dissolution rate of metakaolin comparing to more pronounced at 28 days compared to 7 and 3 days and

Fig. 9. FTIR spectra of geopolymers.

Fig. 10. SEM image of (a) GS22, (b) GS15, (c) GS8 and (d) GS4.

7
A. Dehghani, F. Aslani and N. Ghaebi Panah Construction and Building Materials 293 (2021) 123527

1 day age. For instance, GS22 exhibited a 28-days compressive 9001567), and Quartz (SiO2 – COD ID: 9013321). Phase quantifica-
strength of 37.8 MPa, which was 1.2, 2.8, and 3.4 times larger than tion was carried out using the Rietveld method [63] to investigate
GS15, GS8, and GS4, respectively. The use of a higher content of the geopolymers structure further. Based on the data, amorphous
GGBFS provides more free calcium ions in the system. These ions fractions of all geopolymers developed in this study were between
react with silica and alumina to form calcium silicate hydrate that 81% and 84%, whereas crystalline phases formed only around 20%
co-exists with geopolymeric products, resulting in improved com- of the structure. It can be concluded that although the XRD pat-
pressive strength [49,54,59]. Besides, the heat generated due to the terns of all samples are similar, the chemical bondings between
exothermal reaction of GGBFS and alkaline solutions promotes the atoms in the amorphous area could be different. It is well-
geopolymerisation process [60]. Furthermore, this reaction known that the different arrangement of the chemical bondings
increases the dissolution of fly ash particles by consuming water can modify the chemical network and therefore change the proper-
and increasing the alkalinity of the system, leading to increases ties of the materials. Hence, FTIR analysis was performed to com-
in the rate of polycondensation [54]. The graphs plotted in Fig. 7 pare the chemical bondings formed in geopolymers and potential
also confirm the postulation that the existence of calcium is vital differences among them.
to early-strength development of metakaolin- or fly ash-based Fig. 9 shows FTIR spectra of geopolymers developed in this
geopolymers [18,61,62]. For example, the strength gain after one study. A broad peak in the range of 900–1200 cm1 is attributed
day was only 2.5 MPa and 2.1 MPa for GS8 and GS4, respectively, to Si-O-T (T = tetrahedral Si or Al) [64]. This band is centred at
as the dissolution of silicon and aluminium species from fly ash 978 cm1 in the slag and 1090 cm1 in the fly ash based on their
is very slow in ambient temperature [54]. According to Yip et al. chemical networks [59]. The vibrational band around 960 cm1,
[18], the presence of calcium results in the precipitation of calcium which is the primary asymmetric stretch band of the Si-O-T, iden-
silicate hydrates that provides nucleation sites which then acceler- tified the formation of the geopolymer gel and the incorporation of
ate geopolymerization process. aluminium in the geopolymer structure [9,64]. According to the
FTIR spectra, it seems that the surface of the fly ash particles is
3.4. The structural and morphological characterisation of the mostly reacted with geopolymer as the spectra at 1055 cm1,
geopolymers which is related to the unreacted fly ash, was not clearly visible
[9,64]. The band around 875 cm1 is assigned to the asymmetric
Fig. 8 shows the XRD patterns of the samples with different sil- stretching of AlO4 groups due to the unreacted slag [59]. Compar-
ica/alumina and slag/fly ash ratios listed in Table 2. Comparing the ing the FTIR spectra reveals that the aluminium could bond with
identified phases revealed that there are no significant differences silicon and form geopolymer gel in the samples Si337 (also called
amongst their crystalline structures. The identified phases in all GS22), Si325, Si350, and GS15. Therefore, a relatively high com-
the samples were Corundum (Al2O3 – COD ID: 9008081), Calcite pressive strength is expected for these samples. In contrast, the
(CaCO3 – COD ID: 9000095), Mullite (Al9Si3O19.48 – COD ID: central peak around 972 cm1 in the samples GS8, GS4, and

Fig. 11. SEM image of (a) Si337, (b) Si325, (c) Si350 and (d) Si375.

8
A. Dehghani, F. Aslani and N. Ghaebi Panah Construction and Building Materials 293 (2021) 123527

Fig. 12. Nitrogen sorption isotherms of geopolymers.

Si375 showed that the formation of geopolymer could not create a time and higher flowability have more time to fill porosity and
strong network and some unreacted slag remained, especially in result in a higher bulk density. This can explain the lower BET sur-
the samples of GS8 and GS4. The results of the FTIR were in good face area of the samples with a longer setting time. Therefore, it
agreement with the compressive strength of the samples. can be concluded that the BET surface area in Group 2 is linked
Figs. 10 and 11 show SEM images of the samples with a differ- with the slag content of mixtures. The results listed in Table 3 also
ent initial molar ratio of SiO2/Al2O3 after 28 days. The morphology show the reduction in BET surface area by increasing SiO2/Al2O3
of the samples shows that an integrated geopolymer matrix is ratio (i.e., Si325 to Si375).
formed in all samples. Although some cracks can be seen in
Si337 structure, its compressive strength is the highest value
4. Conclusion
amongst all samples. High compressive strength can be attributed
to a great network bonding and structure of the geopolymer
The following conclusions can be drawn from this study.
matrix. In the case of Si337, the formation of geopolymer provides
The increase in SiO2/Al2O3 ratio leads to shorter initial and final
a dense structure with high compressive strength. These results are
setting times. This observation was quite different from that of
in good agreement with the XRD and FTIR as around 82% of Si337
metakaolin-based geopolymers. Increasing SiO2/Al2O3 ratio also
structure is amorphous in which aluminium ions are bonded with
decreases the flowability of the geopolymer system. Like
silicon ions to form a strong geopolymer gel. In other words, the
metakaolin-based geopolymers, the compressive strength of fly
appropriate initial ratio of SiO2/Al2O3 causes a better
ash-slag geopolymers first increases and then decreases by
geopolymerization.
increasing SiO2/Al2O3 molar ratio. The maximum compressive
Regarding the rate of geopolymer gel formation and setting
strength at different ages (1 day to 91 days) was achieved for a
time, textural properties of the materials could be different.
SiO2/Al2O3 ratio of 3.37 (Si/Al = 1.68), which could be related to
Fig. 12 shows nitrogen sorption isotherms of all samples. All the
the alumina-silica bondings in the amorphous area based on FTIR
curves are identified as type IV isotherms which indicates the pres-
results. This ratio was slightly lower than that reported for
ence of a mesoporous material. Slight differences amongst the hys-
metakaolin-based geopolymers (i.e., Si/Al = 1.9).
teresis loops were also observed. Based on the IUPAC classification
Comparing the identified phases revealed no significant differ-
[65], the hysteresis loops can be categorized into four types from
ences amongst crystalline structures of synthesized geopolymers
H1 to H4, indicating specific pore structures. Cylindrical, bottle-
with different SiO2/Al2O3 molar ratio. Amorphous fractions of all
neck, wedge-shaped, and slit-shaped pores are designated H1,
geopolymers developed in this study were between 81% and 84%,
H2, H3, and H4 pore types, respectively [65].
whereas crystalline phases formed only around 20% of the
BET surface area, pore size, and pore volume of the samples
structure.
were listed in Table 3. Comparing the BET surface area shows a
The increase in fly ash replacement with GGBFS accelerated set-
decreasing trend from Si337 to GS4 samples (i.e., group 2 in
ting times and decreased flowability. It also significantly improved
Table 2), which is consistent with the setting time of samples.
compressive strength, especially at higher ages, due to the forma-
The reduction in the BET surface area was more pronounced when
tion of calcium silicate hydrate co-exists with geopolymeric prod-
the slag content was significantly reduced (GS15 to GS8 and GS4).
ucts. FTIR analysis also showed that geopolymer formation could
The CaO content of the mixture was reduced from Si337 (GS22) to
not create a strong network in the mixtures with low content of
GS4, resulting in a longer setting time. Mixtures with longer setting
GGBFS at ambient curing condition.
Regardless of SiO2/Al2O3 ratio and GGBFS content, all fly ash-
slag geopolymers showed type IV nitrogen sorption isotherms,
Table 3 indicating mesoporous material. Furthermore, it was found that
BET surface area, pore volume, and pore size of geopolymer samples.
BET surface area of geopolymers decreases by increasing SiO2/
Sample BET surface area (m2/g) Pore volume (cm3/g) Pore size (nm) Al2O3 ratio or GGBFS content.
Si325 32.3 0.089 11.0
Si337 32.3 0.076 9.4
Si350 25.3 0.090 14.3
CRediT authorship contribution statement
Si375 24.9 0.079 12.7
GS15 31.3 0.085 10.5 Ayoub Dehghani: Conceptualization, Data curation, Formal
GS8 13.8 0.052 15.0 analysis, Investigation, Methodology, Software, Visualization,
GS4 11.8 0.042 14.3
Writing - original draft. Farhad Aslani: Funding acquisition, Super-
9
A. Dehghani, F. Aslani and N. Ghaebi Panah Construction and Building Materials 293 (2021) 123527

vision, Project administration, Resources, Conceptualization, [18] C.K. Yip, G.C. Lukey, J.L. Provis, J.S.J. van Deventer, Effect of calcium silicate
sources on geopolymerisation, Cem. Concr. Res. 38 (4) (2008) 554–564,
Methodology, Validation, Investigation, Visualization, Writing -
https://doi.org/10.1016/j.cemconres.2007.11.001.
review & editing. Neda Ghaebi Panah: Formal analysis, Writing - [19] E.I. Diaz, E.N. Allouche, S. Eklund, Factors affecting the suitability of fly ash as
review & editing. source material for geopolymers, Fuel 89 (5) (2010) 992–996, https://doi.org/
10.1016/j.fuel.2009.09.012.
[20] M. Zhang, T. El-Korchi, G. Zhang, J. Liang, M. Tao, Synthesis factors affecting
Declaration of Competing Interest mechanical properties, microstructure, and chemical composition of red mud–
fly ash based geopolymers, Fuel 134 (2014) 315–325, https://doi.org/10.1016/
j.fuel.2014.05.058.
The authors declare that they have no known competing finan- [21] J. Davidovits, Geopolymers: inorganic polymeric new materials, J. Therm. Anal.
cial interests or personal relationships that could have appeared 37 (1991) 1633–1656.
to influence the work reported in this paper. [22] J.G.S. van Jaarsveld, J.S.J. van Deventer, G.C. Lukey, The effect of composition
and temperature on the properties of fly ash- and kaolinite-based
geopolymers, Chem. Eng. J. 89 (1-3) (2002) 63–73, https://doi.org/10.1016/
Acknowledgements S1385-8947(02)00025-6.
[23] P.D. Silva, K. Sagoe-Crenstil, V. Sirivivatnanon, Kinetics of geopolymerization:
role of Al2O3 and SiO2, Cem. Concr. Res. 37 (4) (2007) 512–518, https://doi.
Australian Government Research Training Program (RTP) Schol- org/10.1016/j.cemconres.2007.01.003.
arship from The University of Western Australia is gratefully [24] P. Duxson, J.L. Provis, G.C. Lukey, S.W. Mallicoat, W.M. Kriven, J.S.J. van
Deventer, Understanding the relationship between geopolymer composition,
acknowledged. The authors also acknowledge the facilities, and microstructure and mechanical properties, Colloids Surfaces A Physicochem.
the scientific and technical assistance of the Australian Microscopy Eng. Asp. 269 (1-3) (2005) 47–58, https://doi.org/10.1016/
& Microanalysis Research Facility at the Centre for Microscopy, j.colsurfa.2005.06.060.
[25] J. Davidovits, Properties of geopolymer cements, First Int. Conf. Alkaline Cem.
Characterisation & Analysis, The University of Western Australia, Concr. (1994) 131–149.
a facility funded by the University, State and Commonwealth [26] M. Rowles, B. O’Connor, Chemical optimisation of the compressive strength of
Governments. aluminosilicate geopolymers synthesised by sodium silicate activation of
metakaolinite, J. Mater. Chem. 13 (2003) 1161–1165, https://doi.org/10.1039/
b212629j.
References [27] M. Steveson, K. Sagoe-Crentsil, Relationships between composition, structure
and strength of inorganic polymers, J. Mater. Sci. 40 (16) (2005) 4247–4259,
https://doi.org/10.1007/s10853-005-2794-x.
[1] E. Benhelal, G. Zahedi, E. Shamsaei, A. Bahadori, Global strategies and
[28] R.A. Fletcher, K.J.D. MacKenzie, C.L. Nicholson, S. Shimada, The composition
potentials to curb CO2 emissions in cement industry, J. Clean. Prod. 51
range of aluminosilicate geopolymers, J. Eur. Ceram. Soc. 25 (9) (2005) 1471–
(2013) 142–161, https://doi.org/10.1016/j.jclepro.2012.10.049.
1477, https://doi.org/10.1016/j.jeurceramsoc.2004.06.001.
[2] M. Schneider, M. Romer, M. Tschudin, H. Bolio, Sustainable cement
[29] P. De Silva, K. Sagoe-Crenstil, Medium-term phase stability of Na2O–Al2O3–
production—present and future, Cem. Concr. Res. 41 (7) (2011) 642–650,
SiO2–H2O geopolymer systems, Cem. Concr. Res. 38 (6) (2008) 870–876,
https://doi.org/10.1016/j.cemconres.2011.03.019.
https://doi.org/10.1016/j.cemconres.2007.10.003.
[3] D.N. Huntzinger, T.D. Eatmon, A life-cycle assessment of Portland cement
[30] J. Temuujin, A. Minjigmaa, W. Rickard, M. Lee, I. Williams, A. van Riessen,
manufacturing: comparing the traditional process with alternative
Preparation of metakaolin based geopolymer coatings on metal substrates as
technologies, J. Clean. Prod. 17 (2009) 668–675.
thermal barriers, Appl. Clay Sci. 46 (3) (2009) 265–270, https://doi.org/
[4] W. Zhou, C. Yan, P. Duan, Y. Liu, Z. Zhang, X. Qiu, D. Li, A comparative study of
10.1016/j.clay.2009.08.015.
high- and low-Al 2 O 3 fly ash based-geopolymers: the role of mix proportion
[31] J. Davidovits, Mineral polymers and methods of making them, US Patent
factors and curing temperature, Mater. Des. 95 (2016) 63–74, https://doi.org/
4,349,386, 1982.
10.1016/j.matdes.2016.01.084.
[32] D. Hardjito, S.E. Wallah, D.M.J. Sumajouw, B.V. Rangan, On the development of
[5] F. Aslani, A. Deghani, Z. Asif, Development of lightweight rubberized
fly ash-based geopolymer concrete, ACI Mater. J. 101 (2004) 467–472.
geopolymer concrete by using polystyrene and recycled crumb-rubber
[33] S. Greiser, G.J.G. Gluth, P. Sturm, C. Jäger, 29 Si{27 Al}, 27 Al{29 Si} and 27 Al{1
aggregates, J. Mater. Civ. Eng. 32 (2020) 1–16, https://doi.org/10.1061/(ASCE)
H} double-resonance NMR spectroscopy study of cementitious sodium
MT.1943-5533.0003008.
aluminosilicate gels (geopolymers) and gel–zeolite composites, RSC Adv. 8
[6] A.M. Rashad, A comprehensive overview about the influence of different
(70) (2018) 40164–40171, https://doi.org/10.1039/C8RA09246J.
admixtures and additives on the properties of alkali-activated fly ash, Mater.
[34] R.R. Lloyd, J.L. Provis, J.S.J. Van Deventer, Microscopy and microanalysis of
Des. 53 (2014) 1005–1025, https://doi.org/10.1016/j.matdes.2013.07.074.
inorganic polymer cements. 1: remnant fly ash particles, J. Mater. Sci. 44
[7] A.M. Rashad, Alkali-activated metakaolin: a short guide for civil Engineer – an
(2009) 608–619, https://doi.org/10.1007/s10853-008-3077-0.
overview, Constr. Build. Mater. 41 (2013) 751–765, https://doi.org/10.1016/
[35] G. Saini, U. Vattipalli, Assessing properties of alkali activated GGBS based self-
j.conbuildmat.2012.12.030.
compacting geopolymer concrete using nano-silica, Case Stud. Constr. Mater.
[8] V.K. Nagaraj, D.L.V. Babu, Assessing the performance of molarity and alkaline
12 (2020) e00352, https://doi.org/10.1016/j.cscm.2020.e00352.
activator ratio on engineering properties of self-compacting alkaline activated
[36] B.V. Rangan, Fly ash-based geopolymer concrete, Curtin University of
concrete at ambient temperature, J. Build. Eng. 20 (2018) 137–155, https://doi.
Technology, Perth, Australia, Engineering Faculty, 2008. http://hdl.handle.
org/10.1016/j.jobe.2018.07.005.
net/20.500.11937/20680.
[9] K. Juengsuwattananon, F. Winnefeld, P. Chindaprasirt, K. Pimraksa, Correlation
[37] A.M. Mustafa Al Bakri, H. Kamarudin, M. Bnhussain, A.R. Rafiza, Y. Zarina, Effect
between initial SiO2/Al2O3, Na2O/Al2O3, Na2O/SiO2 and H2O/Na2O ratios on
of Na2SiO3/NaOH ratios and NaOH molarities on compressive strength of fly-
phase and microstructure of reaction products of metakaolin-rice husk ash
ash-based geopolymer, ACI Mater. J. 109 (2012) 503–508. 10.14359/51684080.
geopolymer, Constr. Build. Mater. 226 (2019) 406–417, https://doi.org/
[38] M. Olivia, H. Nikraz, Properties of fly ash geopolymer concrete designed by
10.1016/j.conbuildmat.2019.07.146.
Taguchi method, Int. J. Hum. Comput. Interact. 36 (2012) 191–198, https://doi.
[10] D. Khale, R. Chaudhary, Mechanism of geopolymerization and factors
org/10.1016/j.matdes.2011.10.036.
influencing its development: a review, J. Mater. Sci. 42 (2007) 729–746,
[39] P.S. Deb, P. Nath, P.K. Sarker, The effects of ground granulated blast-furnace
https://doi.org/10.1007/s10853-006-0401-4.
slag blending with fly ash and activator content on the workability and
[11] P. Duxson, A. Fernández-Jiménez, J.L. Provis, G.C. Lukey, A. Palomo, J.S.J. van
strength properties of geopolymer concrete cured at ambient temperature,
Deventer, Geopolymer technology: the current state of the art, J. Mater. Sci. 42
Mater. Des. 62 (2014) 32–39, https://doi.org/10.1016/j.matdes.2014.05.001.
(2007) 2917–2933, https://doi.org/10.1007/s10853-006-0637-z.
[40] T. Tho-in, V. Sata, P. Chindaprasirt, C. Jaturapitakkul, Pervious high-calcium fly
[12] C. Ng, U.J. Alengaram, L.S. Wong, K.H. Mo, M.Z. Jumaat, S. Ramesh, A review on
ash geopolymer concrete, Constr. Build. Mater. 30 (2012) 366–371, https://doi.
microstructural study and compressive strength of geopolymer mortar, paste
org/10.1016/j.conbuildmat.2011.12.028.
and concrete, Constr. Build. Mater. 186 (2018) 550–576, https://doi.org/
[41] P. Nath, P.K. Sarker, Use of OPC to improve setting and early strength
10.1016/j.conbuildmat.2018.07.075.
properties of low calcium fly ash geopolymer concrete cured at room
[13] B. Singh, G. Ishwarya, M. Gupta, S.K. Bhattacharyya, Geopolymer concrete: A
temperature, Cem. Concr. Compos. 55 (2015) 205–214, https://doi.org/
review of some recent developments, Constr. Build. Mater. 85 (2015) 78–90,
10.1016/j.cemconcomp.2014.08.008.
https://doi.org/10.1016/j.conbuildmat.2015.03.036.
[42] G. Görhan, G. Kürklü, The influence of the NaOH solution on the properties of the
[14] J. Davidovits, Geopolymers based on natural and synthetic metakaolin-A
fly ash-based geopolymer mortar cured at different temperatures, Compos. Part
critical review, in: 41st Int. Congr. Adv. Ceram. Compos., 2017: pp. 201–214.
B Eng. 58 (2014) 371–377, https://doi.org/10.1016/j.compositesb.2013.10.082.
[15] J.S.J. Provis, J. L. Van Deventer, Introduction to geopolymers, in: J.L. Provis, J.S.J.
[43] J. Davidovits, Geopolymer Cement a review, in: Geopolymer Sci. Tech. Tech.
Van Deventer (Eds.), Geopolymers Struct. Process. Prop. Ind. Appl., Elsevier B.
Pap. No. 21, Geopolymer Institute, 2013: pp. 1–11.
V., 2009: pp. 1–11.
[44] J. Davidovits, Application of Ca-based geopolymer with blast furnace slag, a
[16] J. Davidovits, Geopolymers: inorganic polymeric new materials, J. Therm. Anal.
review, in: P.T. Jones, Y. Pontikes, J. Elsen (Eds.), Second Int, Katholieke
Calorim. 37 (8) (1991) 1633–1656.
Universitat, Leuven, Belgium, Slag Valoris. Symp. Transit. to Sustain. Mater.
[17] X. Hua, J.S.J. Van Deventer, Geopolymerisation of multiple minerals, Miner.
Manag., 2011, pp. 33–49.
Eng. 15 (2002) 1131–1139, https://doi.org/10.1016/S0892-6875(02)00255-8.

10
A. Dehghani, F. Aslani and N. Ghaebi Panah Construction and Building Materials 293 (2021) 123527

[45] AS/NZS (Australian/New Zealand Standards), Supplementary cementitious [56] L. Mo, L. Lv, M. Deng, J. Qian, Influence of fly ash and metakaolin on the
materials, Part 1: Fly ash, AS/NZS 3582.1, Sydney, Australia; Wellington, New microstructure and compressive strength of magnesium potassium phosphate
Zealand, 2016. cement paste, Cem. Concr. Res. 111 (2018) 116–129, https://doi.org/10.1016/j.
[46] AS 3582.2, Supplementary cementitious materials Part 2: Slag—Ground cemconres.2018.06.003.
granulated blast furnace, Standard Australia, 2016. [57] Ch. Panagiotopoulou, E. Kontori, Th. Perraki, G. Kakali, Dissolution of
[47] ASTM, C191, Standard Test Method for Time of Setting of Hydraulic Cement by aluminosilicate minerals and by-products in alkaline media, J. Mater. Sci. 42
Vicat Needle, ASTM C191, West Conshohocken, PA, United States (2019), (9) (2007) 2967–2973, https://doi.org/10.1007/s10853-006-0531-8.
https://doi.org/10.1520/C0191-13.2. [58] J.L. Provis, J.S.J. van Deventer, Geopolymerisation kinetics. 2. Reaction kinetic
[48] Prinya Chindaprasirt, Pre De Silva, Kwesi Sagoe-Crentsil, Sakonwan modelling, Chem. Eng. Sci. 62 (9) (2007) 2318–2329, https://doi.org/10.1016/j.
Hanjitsuwan, Effect of SiO2 and Al2O3 on the setting and hardening of high ces.2007.01.028.
calcium fly ash-based geopolymer systems, J. Mater. Sci. 47 (12) (2012) 4876– [59] I. Ismail, S.A. Bernal, J.L. Provis, R. San Nicolas, S. Hamdan, J.S.J. van Deventer,
4883, https://doi.org/10.1007/s10853-012-6353-y. Modification of phase evolution in alkali-activated blast furnace slag by the
[49] S. Kumar, R. Kumar, S.P. Mehrotra, Influence of granulated blast furnace slag on incorporation of fly ash, Cem. Concr. Compos. 45 (2014) 125–135, https://doi.
the reaction, structure and properties of fly ash based geopolymer, J. Mater. org/10.1016/j.cemconcomp.2013.09.006.
Sci. 45 (2010) 607–615. 10.1007/s10853-009-3934-5. [60] T. Phoo-ngernkham, A. Maegawa, N. Mishima, S. Hatanaka, P. Chindaprasirt,
[50] S. Antiohos, S. Tsimas, Activation of fly ash cementitious systems in the Effects of sodium hydroxide and sodium silicate solutions on compressive and
presence of quicklime: Part I. Compressive strength and pozzolanic reaction shear bond strengths of FA–GBFS geopolymer, Constr. Build. Mater. 91 (2015)
rate, Cem. Concr. Res. 34 (2004) 769–779, https://doi.org/10.1016/j. 1–8, https://doi.org/10.1016/j.conbuildmat.2015.05.001.
cemconres.2003.08.008. [61] R.R. Lloyd, J.L. Provis, J.S.J. van Deventer, Microscopy and microanalysis of
[51] K. Dombrowski, A. Buchwald, M. Weil, The influence of calcium content on the inorganic polymer cements. 2: the gel binder, J. Mater. Sci. 44 (2009) 620–631.
structure and thermal performance of fly ash based geopolymers, J. Mater. Sci. 10.1007/s10853-008-3078-z.
42 (9) (2007) 3033–3043, https://doi.org/10.1007/s10853-006-0532-7. [62] J. Temuujin, A. van Riessen, R. Williams, Influence of calcium compounds on
[52] P. Nath, P.K. Sarker, Effect of GGBFS on setting, workability and early strength the mechanical properties of fly ash geopolymer pastes, J. Hazard. Mater. 167
properties of fly ash geopolymer concrete cured in ambient condition, Constr. (1-3) (2009) 82–88, https://doi.org/10.1016/j.jhazmat.2008.12.121.
Build. Mater. 66 (2014) 163–171, https://doi.org/10.1016/ [63] D.L. Bish, S.A. Howard, Quantitative phase analysis using the Rietveld method,
j.conbuildmat.2014.05.080. J. Appl. Crystallogr. 21 (1988) 86–91. 10.1107/S00218898870094
[53] M.H. Al-Majidi, A. Lampropoulos, A. Cundy, S. Meikle, Development of [64] C.A. Rees, J.L. Provis, G.C. Lukey, J.S.J. van Deventer, In situ ATR-FTIR study of
geopolymer mortar under ambient temperature for in situ applications, the early stages of fly Ash geopolymer gel formation, Langmuir 23 (2007)
Constr. Build. Mater. 120 (2016) 198–211, https://doi.org/10.1016/ 9076–9082. 10.1021/la701185g.
j.conbuildmat.2016.05.085. [65] K.S.W. Sing, D.H. Everett, R.A.W. Hall, L. Moscou, R.A. Pierotti, J. Rouquerol, T.
[54] S. Puligilla, P. Mondal, Role of slag in microstructural development and Siemieniewska, Reporting physisorption data for gas/solid systems with
hardening of fly ash-slag geopolymer, Cem. Concr. Res. 43 (2013) 70–80, special reference to the determination of surface area and porosity, Pure
https://doi.org/10.1016/j.cemconres.2012.10.004. Appl. Chem. 57 (1985) 603–619.
[55] O. Boukendakdji, S. Kenai, E.H. Kadri, F. Rouis, Effect of slag on the rheology of
fresh self-compacted concrete, Constr. Build. Mater. 23 (7) (2009) 2593–2598,
https://doi.org/10.1016/j.conbuildmat.2009.02.029.

11

You might also like