1 s2.0 S0950061819330260 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Construction and Building Materials 236 (2020) 117574

Contents lists available at ScienceDirect

Construction and Building Materials


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

Effects of ceramic tile powder waste on properties of self-compacted


alkali-activated concrete
Ghasan Fahim Huseien a,b, Abdul Rahman Mohd Sam a, Kwok Wei Shah b,⇑, Jahangir Mirza c,⇑
a
Faculty of Engineering, Universiti Teknologi Malaysia, Skudai 81310, Malaysia
b
School of Design and Environment, National University of Singapore, Singapore 117566, Singapore
c
School of Engineering, University of Guelph, Guelph, Ontario N1G 2W1, Canada

h i g h l i g h t s

 Reports properties of self-compacting alkali-activated concrete incorporating wastes.


 Increasing ceramic powder waste enhances the workability performance.
 Inclusion of ceramic tile powder waste lowers the concrete strength.
 Enhancement of concrete durability with the addition of ceramic wastes.
 Achieved low carbon footprint by using high volume ceramic powder content.

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

Article history: This study explored rheological and hardened-state properties of self-compacting alkali-activated
Received 1 June 2019 concrete (SCAAC) incorporating ceramic tile powder waste (CPW) for construction activities. Samples
Received in revised form 22 October 2019 containing 100% ground blast furnace slag (GBFS) were used as reference. Test data revealed that the
Accepted 9 November 2019
mini-slump flow enhanced, while the plastic viscosity of SCAAC reduced when 50% or higher concentra-
Available online 19 November 2019
tion of CPW was mixed with GBFS. Also, CPW affected the workability and plastic viscosity considerably
when used in higher concentration. The strength decreased as GBFS substitution with CPW increased. In
Keywords:
conclusion, the chemical structure, capacity and water assimilation of GBFS underpinned the rheological
Self-compacting alkali-activated concrete
Ceramic wastes
behaviour of SCCAC.
GBFS Ó 2019 Elsevier Ltd. All rights reserved.
Acid resistance

1. Introduction egy to attenuate the negative environmental effects of cement pro-


duction is to use industrial and agriculture wastes by partially
Self-compacting concrete constitutes a major advancement in substituting the cement [12–14]. Another strategy proposed
construction industry. It is characterised by flow-ability and con- recently is the use of geopolymer as a new constructional binder,
solidation under its own weight, permeate the formwork fully by because the CO2 emissions it generates are far lower than those
penetrating the spaces separating the reinforcement bars and keep associated with standard Portland cement [15–17]. Production of
its stable composition at the same time [1–3]. Self-compacting aluminosilicate-based inorganic polymers, known as geopolymer,
concrete has many advantages over conventional concrete, includ- are achieved through synthesis of waste pozzolanic materials with
ing decreased construction time and labor cost, reduced noise pol- highly alkaline hydroxide and/or alkaline silicate [18,19]. Geopoly-
lution, improved filling capacity of highly congested structural merisation can be achieved by using industrial solid wastes and
members, enhanced interfacial transitional zone between cement by-products enriched in silica and/or alumina contents (e.g. fly
paste and aggregate or reinforcement, decreased permeability ash) as pozzolanic components [20–24].
and improved durability of concrete [4–6]. In construction indus- Alkali-activated mortars and concrete have started to distin-
try, cement production alone is responsible for about 7% of the guish from standard Portland cement as an effective constructional
overall amount of CO2 emissions at global level [7–11]. One strat- binder without high CO2 emissions [25–27]. An inorganic polymer
material, alkali-activated concrete consists of calcium (Ca) and
⇑ Corresponding authors. aluminosilicates (ASs). It is synthesised from alkaline activator
E-mail addresses: bdgskw@nus.edu.sg (K.W. Shah), mirzaj@uoguelph.ca solution containing sodium silicate (NS) and sodium hydroxide
(J. Mirza).

https://doi.org/10.1016/j.conbuildmat.2019.117574
0950-0618/Ó 2019 Elsevier Ltd. All rights reserved.
2 G.F. Huseien et al. / Construction and Building Materials 236 (2020) 117574

(NH) and pozzolanic compounds [28,29]. The production of alkali- alkalines unit and gel of calcium silicate hydrate (C-S-H). More-
activated concrete/mortars is based on the exploitation of a wide over, evidence has also been put forward that GBFS made Ca pro-
range of by-products and agricultural wastes comprising calcium, gressively more soluble and improved products’ inhomogeneity
silica and aluminium (Al) such as, GBFS, palm oil fuel ash (POFA) and amorphosity. Additionally, concrete compressive strength is
and fly ash (FA) with low and high calcium, etc., [30,31]. improved by the development of N(C)-A-H C-S-H and products of
Besides being non-recyclable, ceramic waste is also bulky and Na/Ca -aluminosilicate-hydrate [10,51–55].
creates problems for landfill disposal. Hence, for the purposes of Self-compacting green concrete and high-performance concrete
natural resource conservation and environmental protection, it is have led to the development of a new type of cement-free concrete,
necessary to develop new products with ceramic waste to recycle known as self-compacting high-performance alkali-activated con-
and to use them in construction projects. The pozzolanic properties crete (SCAAC). This type of concrete displays not only self-
of ceramic waste have been confirmed by earlier studies [32–34]. compacting concrete properties, such as filling and penetration
Because they benefit the mechanical and durability properties of capability as well as ability to withstand separation, but is also
concrete, natural pozzolanic wastes were employed as construc- extremely strong and durable [3]. However, the mechanical prop-
tion material in earlier times [35,36]. However, industrial wastes erties of SCAAC on its own and in combination with ceramic waste
are currently the preferred pozzolanic materials because of strict have not been extensively studied. Therefore, to fill this gap in the
environmental policies [37–39]. The ceramic industry has a signif- literature, the present study used different concentrations of CPW
icant effect on the environment and relies on landfills to dispose of to substitute GBFS as cement-free binder, with minimal NH molar-
the massive volumes of its waste. In 2015, ceramic tile production ity and low viscosity of alkaline activator solution comprising
at global level was around 12.4 million square metres [40,41], of sodium silicate (NS) and NH. The present study mainly aims to
which 10–30% is discarded as waste [32,42]. Additionally, the decrease the solid wastes’ amount in the environment, thus
greatest proportion of this waste is non-recyclable, which is highly lowering the CO2 emissions. The other aims are to suggest feasible
problematic in terms of disposal. To address this issue of extensive materials and physical requirements to produce alkali-activated
amounts of ceramic waste with little usage, researchers started GBFS-CPW self-compacting concrete, as well as to propose
investigating its use in mortar or concrete instead of cement. This, alternative materials for use in the construction industry.
in turn, is advantageous as it reduces cost and energy consump-
tion, improves ecological equilibrium and promotes more careful 2. Experimental program
use of natural resources [43–46].
Due to highly concentration of calcium and silica, the GBFS car- 2.1. Materials
ries both the pozzolanic and cementing qualities, which makes it a
popular material in the construction industry for making modified Binders, aggregate and alkaline-activated solution were the
concrete more durable with enhanced mechanical properties [47]. three types of materials employed in the present study. CPW and
It was reported that GBFS changed the microstructure and durabil- GBFS were the waste materials used as binders in the production
ity of alkali-activated concrete to which it was added [23,48]. of SCAACs. Pure GBFS (Ipoh, Malaysia) was a resource material that
Meanwhile, Yusuf et al. [49,50] observed that alkali-activated mor- was employed to produce cement-free binder and its subsequent
tar was stronger when GBFS was added to it as binder. This was use did not require laboratory treatment. GBFS is characterised
explained in terms of the fact that GBFS served to fill pores and by its cementing and pozzolanic qualities. It is off-white in colour
led to the development of uniform microstructures with high and results in a hydraulic reaction with water. With a composition
arrangement and twin products comprising highly polymerised of calcium silicate and alumina (around 90%), GBFS satisfies the

Fig. 1. Processing steps of CPW.


G.F. Huseien et al. / Construction and Building Materials 236 (2020) 117574 3

Table 1
Chemical composition and physical characteristics of CPW and GBFS.

Chemical composition (% by mass)


Oxide CaO Al2O3 SiO2 Fe2O3 MgO K2O Na2O SO3 LOI
CWP 0.02 12.6 72.6 0.56 0.99 0.03 13.5 0.01 0.13
GBFS 51.8 10.9 30.8 0.64 4.57 0.36 0.45 0.06 0.22
Physical characteristics
Item Specific gravity Medium particle size, mm Specific surface area, m2/g
CWP 2.6 35 12.2
GBFS 35 12.8 13.6

Fig. 2. Particle size distribution of CPW and GBFS.


Fig. 3. XRD patterns of GBFS and CPW.

pozzolanic material specification imposed by ASTM C618. The con-


struction industry in Johor, Malaysia, supplied the ceramic waste the extreme amorphous characteristic of this material was con-
employed in the current study. The first step upon acquisition of firmed by the XRD pattern, which did not show any sharp peak.
this material was to separate large particles by crushing followed The reactive silica and calcium contents are the main contributors
by sieving through 600 lm mesh. After sieving, the material was to the development of GBFS. In fact, the possibility of alkali-
ground for six hours, with hourly inspection to examine how fine activated preparation is due to high levels of reactive amorphous
the particles were. The subsequent step was to collect and to use silica and calcium in GBFS. Fig. 4 illustrates the images of CPW
the powder in the process of mixing with water. Fig. 1 illustrates and GBFS produced through scanning electron microscopy (SEM).
the manner in which the ceramic waste was treated. It can be clearly seen that both materials contained irregular and
To determine how GBFS and CPW were chemically constituted, angular shaped particles, corroborating earlier studies [20,58].
X-Ray fluorescence (XRF) spectroscopy was applied (Table 1). Fourier transform infrared (FTIR) analysis provides detailed sig-
Results showed that GBFS and CPW primarily consisted of silica nature of chemical bonding vibrations of the constituents and
and aluminium in proportion of 41.7% and 84.8%, respectively. Fur- allows in making a comprehensive structural analysis. Transforma-
thermore, the concentration of calcium oxide was significantly tion of GBFS and CPW from crystalline to amorphous state with
higher in GBFS (51.8%) than in CPW. It is well known that the syn- temperature was examined by recording their FTIR spectra in
thesis of alkali-activated mortars/concrete depends greatly on the terms of wavenumber (cm1) which is equivalent to the frequency
concentration of silica, aluminium and calcium oxide. Meanwhile, or energy of bonding vibrations (Fig. 5). The stretching modes are
neither GBFS nor CPW had a high concentration of potassium oxide sensitive to the Si–Al composition of the framework and may shift
(K2O). Additionally, both GBFS and CPW conformed to ASTM C618 to a lower frequency with increasing number of tetrahedral alu-
standard as they had extremely low loss on ignition contents. By minium atoms [59]. Sitarz and Handke [60] reported the occur-
using the particle size analyser test, the particle size distributions rences of broad band at 500–650 cm1, indicating the existence
of both GBFS and CPW were calculated and demonstrated in of silicate and aluminosilicate glass phases. It also possessed
Fig. 2. The median of particles for GBFS and CPW were estimated short-range structural order in the form of tetrahedral or octahe-
to be 12.8 and 35 mm, respectively. For physical properties of raw dral rings. The spectral bands at 460 cm1 was related to Al–O/
materials including GBFS and CPW, the surface areas were evalu- Si–O in plane and bending modes, 730 cm1 was assigned to octa-
ated and presented in Table 1. hedral site modes of Al, 820 cm1 was allocated to tetrahedral Al–O
The results of X-Ray Diffraction Patterns (XRD) of GBFS and stretching [61] and, 1400 cm1 was endorsed to asymmetric
CPW are shown in Fig. 3. For CPW, the crystalline silica and alu- stretching vibrations of Al–O/Si–O bonds [62]. The CPW presented
mina compounds were believed to be the reason for the marked a broad band with medium intensity (C=O) between (1460–1600)
diffraction peaks exhibited by the XRD pattern at about 2h = 16– cm1. Furthermore, CPW displayed the highest intensity compared
30° [17,56,57], whereas the crystalline quartz and mullite phases to GBFS. The GBFS did not show any vibrational band in this region.
were identified as the cause of other crystalline peaks. For GBFS, In both the waste materials, the bands appeared in the regions of
4 G.F. Huseien et al. / Construction and Building Materials 236 (2020) 117574

Fig. 4. Scanning electron microscopy (SEM) images of CPW and GBFS.

Fig. 5. FTIR spectra of CPW and GBFS.

3500 cm1 were assigned to bending vibrations of (C=O-H) and friendlier to environment and is low in cost, energy consumption
stretching vibration of (–OH), respectively. Hydroxyl ions being and carbon dioxide emission. Sodium hydroxide solution with
characteristic of weakly bound ligament of water were either 2 M was prepared by using analytical grade NaOH (NH, 98% purity)
adsorbed on the surface or trapped in large cavities. in the form of pellets; then the NH pellets were disbanded in water.
Table 2 shows the composition of alkaline activator solution The prepared solution of sodium hydroxide was left for 1 day to
used in this study. To reduce the environmental effects of NaOH cool down and then added up to a sodium silicate (NS) solution
and Na2SiO3, low molarity (2 M) of NaOH solution and Na2SiO3 to to achieve the final activator solution having SiO2:Na2O ratio of
NaOH ratio of 0.75 were adopted to prepare the alkaline activator 1.2. Analytical grade NS solution comprising of Na2O (14.70 wt%),
solution. These were fixed for all solution dosages. Regarding the SiO2 (29.5 wt%) and H2O (55.80 wt%) was utilized. The ratio of NS
contents of Na2O, SiO2 and H2O, the prepared alkaline solution is to NH was fixed to 0.75 for all alkaline activator solutions.

Table 2
Composition of alkaline activator solutions.

NaOH solution (NH) Na2SiO3 solution (NS) NS:NH Final of Ms


Mass% (SiO2:Na2O)
Molarity M Na2O H2O SiO2 Na2O H2O SiO2:Na2OMass%
Mass%
Mass% Mass% Mass% Mass% Mass%
2 7.4 92.6 29.5 14.7 55.8 2.01 0.75 1.2
G.F. Huseien et al. / Construction and Building Materials 236 (2020) 117574 5

Table 3 3. Testing of fresh and hardened SCAACs


Physical characteristics of fine and coarse aggregates.

Material Water absorption Specific Maximum size 3.1. Workability tests


(%) gravity (mm)
Fine aggregate 1.2 2.62 2.36 Since a standard technique is yet to be devised to evaluate all
Coarse aggregate 0.51 2.67 10 three workability properties of SCAACs, multiple techniques were
employed in this study to investigate those properties for every
mix sample. Table 5 provides the techniques to assess the filling
and penetration capabilities, while observation through visual sta-
The physical properties of fine and coarse aggregates used to bility is indicative of separation resistance. A range of test methods
prepare SCAACs mixtures are presented in Table 3. River sand for evaluation of SCAAC mixes have been put forth by the European
was obtained from a local supplier. The specific gravity of 2.62 Guidelines EFNARC. According to these guidelines, this study
was used as fine aggregate under saturated surface dry condition assessed the workability properties based on the methods of slump
for all mixes to guarantee that alkaline solution binder (S:B) ratio flow, V-funnel, L-Box, T-50 and J-Ring. In terms of alkali-activated
remains unaffected. For coarse aggregate, the crushed granite, materials setting time, a distinction was made between initial and
obtained from the quarry, was used for all the mixes with specific final setting time. Mortar setting time was measured with the Vicat
gravity 2.67 and water absorption 0.51%. method, in keeping with the requirements of ASTM C191.

2.2. Mix design 3.2. Hardened concrete testing

CPW and GBFS were mixed in different concentrations to pro- According to ASTM C109-109 M, the compressive strength test
duce SCAACs. More specifically, the concentration of CPW ranged was conducted on samples after curing them in air properly for a
from 10 to 80% by weight, while 100% GBFS served as control period of 3, 7, 28 and 56 days. After every curing interval, testing
and its concentration ranged from 20 to 100% by weight. This was conducted on three specimens at each age, with every speci-
binary-blend in different concentrations was employed to con- men being adequately prepared and placed between the superior
duct several tests to achieve the research aim of determining and inferior metal bearing plates. The specimens were subjected
how SCAACs were affected by high CPW concentrations. All to a fixed 2.5 kN/s loading rate until failure. Calculation of com-
other compounds used in the mixtures did not vary in concen- pressive strength was based on standard deviation of 4.3 MPa
tration as CPW and GBFS were. Table 4 provides an overview and coefficient of variation of 5.4. Assessment of tensile strength
of SCAAC mixtures. NH and NS were blended rigorously for the and flexural strength were carried out following ASTM C496/
preparation of a mixture measured by weight. This process C496M-17 and ASTM C78, respectively. Both assessments were
released heat, so the mixture was left to reach room tempera- conducted after the specimens were cured for 28 days. The results
ture prior to being used. For the preparation of SCAACs, a con- were outcome of the average of three samples. Furthermore, com-
crete mixer machine was employed to blend CPW and GBFS parison between the measured strengths and the control sample of
for a period of three minutes under dry conditions until a uni- 100% GBFS was also undertaken.
form mixture of fine and coarse aggregates was obtained. This To assess how porous the generated SCAACs were, the vacuum
was followed by the addition of the alkaline solution to the mix- saturation technique was applied as per ASTM C642-13. The satu-
ture, which was subjected to mixing for a further six minutes in rated, suspended and oven-dry masses tests provided the required
the concrete mixer machine at moderate speed. After that, a parameters. These tests were conducted on specimens cured for
number of techniques were applied to assess the prepared 28 days with saturated 100 mm  100 mm surface. After breaking
SCAACs in terms of main properties that SCAACs must fulfil to the alkali-activated specimens at 28 days, their middle portions
be classified as concrete mixtures, namely, filling and penetra- were ground into a fine powder and subjected to microstructural
tion capability as well as separation resistance. assessment via XRD, SEM and FTIR.

Table 4
Self-compacting alkali-activated concrete mix design.

Mix code Binder kg/m3 Aggregate kg/m3 Alkaline activator solution kg/m3
GBFS CPW Sand Coarse S:B M NS:NH NS NH
SCAAC1 484 0
SCAAC2 338.8 145.2
SCAAC3 290.4 193.6 844 756 0.50 2 0.75 104 138
SCAAC4 242 242
SCAAC5 193.6 290.4
SCAAC6 145.2 338.8

S:B, Ratio of alkaline activator solution to binder; M, molarity of sodium hydroxide; NS:NH, Sodium silicate to sodium hydroxide; NS, sodium silicate; NH, sodium hydroxide.

Table 5
Self-compacting concrete acceptance criteria as per EFNARC [63].

Test Slump flow, mm T50 flow, sec V-funnel, sec L-box ratio, H2:H1 J-ring, mm
Acceptance criteria as per EFNARC
Min 650 2 6 0.8 0
Max 800 5 12 1.0 10
6 G.F. Huseien et al. / Construction and Building Materials 236 (2020) 117574

Fig. 6. Slump flow of SCAACs containing different ratios of CPW replacing GBFS.
Fig. 7. T50 flow of SCAACs containing different ratios of CPW replacing GBFS.
3.3. Sulphuric acid attack test

The sulphuric acid solution (H2SO4) mainly attacks SCAACs by


dissolving the binder paste matrix, leading to weakening of
strength of effected concrete. The effect of sulphuric acid on con-
crete compositions under study is the 10% H2SO4 acid solution
which is prepared with deionised water. For each concrete mixture
and after 28 days of curing age, six specimens were selected and
weighed before being immersed in H2SO4 solution for one year.
Throughout the experiment and every three months, the H2SO4
solution was changed, in order to maintain a continuity of acid
solution’s pH level. After 12 months, the SCAACs’ specimens were
checked while several different factors were considered during
evaluating their performance, including remaining strength, loss
in weight and ultrasonic pulse velocity according to ASTM C267
specifications [64].

4. Results and discussion

4.1. Filling ability of concrete


Fig. 8. H2:H1 ratio of SCAACs containing different ratios of CPW replacing GBFS.

Three tests, namely slump flow, T50 and L-box tests were
water demand for CPW compared to GBFS [23,36,65]. As reported
adopted in this study to measure the impact of CPW contents on
in previous studies [47,66], the low water demand affects the
filling ability of concrete. The slump flow test results (Fig. 6)
chemical reaction rate and increases the plasticity of mixture
showed a significant influence on the concrete workability with
which improves the concretes’ workability.
increasing replacement level of GBFS by CPW in alkali-activated
matrix. With increasing CPW content replacing GBFS from 0 to
80%; the flow ability enhanced and its flow diameter increased 4.2. Passing ability of concrete
from 560 mm to 780 mm, respectively. Concrete mixes containing
50, 60, 70 and 80% of CPW as GBFS replacement, achieved high Fig. 9 shows the effect of CPW content on J-ring test results of
workability and are considered as self-compacting mixes according concrete formulated from SCAACs at different CPW:GBFS ratios.
to EFNARC [63]. Fig. 7 shows the effects of CPW levels on time flow It was observed that all mixtures containing 30% CPW and up
(T50) values of concrete mixes. The results showed a great enhance- achieved the criteria requirement of self-compacting concrete. As
ment on the workability of concrete with increased level of CPW the CPW contents increased from 0 to 10, 20, 30, 40, 50, 60, 70
replacing GBFS. The time value dropped from 6 s to 2.5 s with and 80% to replace GBFS, the workability of concrete slightly tends
decreasing contents of GBFS as the replacement level with CPW to enhance and achieved the requirements. The highest value was
increased. Most mixes prepared with 40% CPW replacing GBFS or obtained with mixture containing 80% CPW which presented
more achieved the EFNARC requirements for self-compacting con- 4.5 mm compare to 12 mm achieved with control sample (100%
crete. The last test conducted to evaluate the filling ability of con- GBFS). Several researchers [36,46,57] reported a reduction in CaO
crete was L-box. Its results revealed that all the samples passed the content which led to enhance the workability due to a slow poz-
requirement for SCC, except for samples containing 0 and 10% CPW zolanic chemical reaction.
which indicated the ratio of 0.79 and 0.77, respectively (Fig. 8). The
increments in flow ability of concrete could majorly be influenced 4.3. Resistance to segregation
by large particles size distribution and low specific surface area of
CPW compared to GBFS (as presented in Table 1). Furthermore, the The resistance to segregation of SCAAC mixes were evaluated
enhancement in concretes workability could be attributed to low using V-funnel test. The results showed a high influence on the
G.F. Huseien et al. / Construction and Building Materials 236 (2020) 117574 7

Fig. 9. J-Ring flow of SCAACs containing different ratios of CPW replacing GBFS. Fig. 11. Setting times of SCAACs containing different ratios of CPW replacing GBFS.

workability of concrete by the CPW level in alkali-activated matrix CPW contents present in alkali-activated matrix. The highest initial
(Fig. 10). Compare to control sample and criteria for self- and final setting times recorded were 64 min and 108 min, respec-
compacting mix, all the mixtures containing 30%-70% CPW as tively in SCAAC mix containing 80% CPW. This behaviour could be
replacement of GBFS achieved the requirements. The increment attributed by the lower alkali activation due to higher amounts of
in workability of concrete including filling-ability, passing and silica present in the higher CPW contents in SCAACs which resulted
resistance to segregation could be attributed due to reduced CaO in lower setting time than with 100% GBFS. The difference in vari-
content while increasing CPW level replacing GBFS. The rapid rate ation between the initial and final setting times is low. However,
of chemical reaction of high volume GBFS content decreases the this difference in both the setting times increased with the increas-
plasticity of mixture which is responsible in the reduction of con- ing CPW contents in the matrix. This is due to slow pozzolanic
crete’s workability [57]. Also the low surface area of CPW com- reaction between SiO2 (from CPW) and Ca(OH)2 evolved from GBFS
pared to GBFS reduced the water demand from alkaline solution hydration which only occurs after the latter started to react with
and led to enhance the workability of alkali-activated matrixes alkaline activator solution and water [36,46,47,68].
[67]. In the mixture of SCAACs6, the increased content of CWP to 80%
and decreased content of GBFS to 20% led to decrease the CaO
amount and increase the SiO2 content which resulted in a delayed
4.4. Setting time chemical reaction rate and increased setting time. It also supported
the fact that higher the CaO content in the alkali-activated matrix,
Fig. 11 illustrated the results of initial and final setting times of the faster is the setting rate [36,69]. These results confirmed that
SCAACs prepared with varying levels of CPW replacing GBFS as the CPW, as a part of the binary blended alkali-activated binder,
binder. It showed that the initial and final setting times ranged was effective to decelerate the setting time of concretes at ambient
from 6 to 64 min and 10 to 108 min, respectively. The control conditions.
sample containing 100% GBFS recorded the lowest value for both
the initial and final setting times. The initial and final setting times
of alkali-activated pastes increased slightly with the increase in 4.5. Compressive strength

Fig. 12 summarizes the compressive strength test results of


SCAACs. Each presented value consisted of average of three mea-
surements. The effects of CPW including alkali-activated matrix
as GBFS replacement were evaluated and discussed in this section.
Various levels of CPW replacing GBFS starting from 10 to 80% by
mass were assessed and compared with control sample containing
100% GBFS. It is evident from Fig. 12 that the use of CPW as a GBFS
replacement decreased the early strength of the SCAAC mixes com-
pared to the control at all the test ages. At age of 3 days, the control
sample (with 100% GBFS) provided the highest value of compres-
sion strength, i.e., 52.6 MPa. As soon as GBFS was replaced by
10% CPW in SCAACs, the compressive strength dropped to
47 MPa. This strength loss continued with the increased level of
CPW from 10 to 20, 30, 40, 50, 60, 70 and 80%. The lowest compres-
sive strength value of 18.6 MPa was recorded in a mixture contain-
ing 80% of CPW at age of 3 days. However, these results indicated
that the development in compressive strength was monotonically
increased with the increase in age. After 28 days of age, the com-
pressive strength gain rate continued to increase for all SCAACs
Fig. 10. V-funnel of SCAACs containing different ratios of CPW replacing GBFS. mixtures and showed a percentage increase higher than 32%. At
8 G.F. Huseien et al. / Construction and Building Materials 236 (2020) 117574

in alkali-activated matrix. The lowering in compressive strength


could be attributed to several factors. The 1st factor is the differ-
ence in chemical composition of raw materials used (CPW and
GBFS) which affected significantly the alkali-activation process of
the binders. The 2nd factor effects on compressive strength devel-
opment is the lower reaction rates of CPW comparing to GBFS
which was partially dissolved [70]. The 3rd factor could be associ-
ated with the lowering of compactness and density of concrete
specimens with rising CPW amounts. The 4th factor is linked to
the low NaOH concentration (2 M) where the compressive strength
depended mainly on the level of CaO to replace the low amount of
Na2O. Thus, the formation of more C-S-H and C-A-S-H gels beside
the N-A-S-H gel improved the prepared concretes’ compressive
strength develeopment [17,57].

4.6. X-ray Diffraction analysis

The X-ray Diffraction patterns (XRD) of SCAACs in which GBFS


Fig. 12. Effect of CPW replacing GBFS on compressive strength of SCAACs. was substituted with CPW to various degrees are illustrated in
Fig. 13. It is important to mention here that glass content of fly
ash consists of reactive and refractory glass, and only the former
age of 56 days, the compressive strength increased between 3 and participates in geopolymerisation. The broad and diffused back-
16% with increasing hydration time as compared to specimens ground peak with maxima of around 10° was emerged from the
tested at 28 days. The reduction in compressive strength of con- short-range order of the hydrotalcite (Mg6Al2CO3OH164H2O)
crete could be due to reduced CaO content which led to increase [69,71]. Furthermore, the high content of MgO (4.56%) in GBFS
in silicate to calcium ratio (SiO2:CaO) and produced a negative (Table 1) formulated this peak during the geopolymerization pro-
effect on compressive strength of SCAACs. Moreover, the produc- cess. An increase in CPW concentration showed a reduction in C-
tion of C-(A)-S-H gels were significantly influenced by increased S-H peak intensity, particularly the 30° peak. At 70% level of
CPW contents and reduction in CaO contents which resulted CPW, a hydrotalcite peak substituted the C-S-H peak at 43°. Mean-
slower chemical reaction rates and lower strength. The occurrence while, at CPW level higher than 50%, a quartz peak substituted the
of low amount of C-(A)-S-H gels could be ascribed to low Ca con- calcite peak at 17°, 21°, 28°, 38.5° and 51°. Moreover, the non-
tent which played a major role in loss of concretes strength. In a reacted silicate quantity increased and the C-S-H product
previous study which was reported by Rashad [26] who observed decreased as the CPW level increased, which was due to the fact
a decrease in compressive strengths with increasing content of that CS diminished from 52.8 to 34.1 MPa (at age of 28 days). In
materials containing high amount of ASs, such as POFA and FA, previous studies [58,72], it was reported that a reduction in CaO

Fig. 13. Effect of CPW replacing GBFS on XRD results of SCAACs.


G.F. Huseien et al. / Construction and Building Materials 236 (2020) 117574 9

Fig. 14. Effects of CPW replacing GBFS on microstructure of SCAACs (a) 50% CPW (b) 70% CPW.

amount effects to produce less C-S-H gel and restricted the calcite 4.8. Fourier transform infrared spectroscopy (FTIR)
formulation. Furthermore, the N-A-S-H gel was found to co-exist
with C-S-H gel and calcite [73]. The compressive strength outcomes indicated that specimens
became weaker as the concentration of CPW substituting GBFS
increased in the SCAAC matrix. As the concentration of GBFS
4.7. Scanning electron microscopy (SEM) declined, so did the levels of dissolved Ca and Al, which was the
reason for the gradual reduction in strength. Furthermore, an ear-
SEM images of SCAACs incorporating CPW in various concentra- lier study observed that the extent of silicate polymerisation was
tions are shown in Fig. 14. Examination at age of 28 days was also diminished by the decrease in dissolved Al [23]. An overview
focused on two specimens with CPW substituting GBFS in propor- of the FTIR spectral analysis is provided in Fig. 15. This analysis
tion of 50% and 70%, respectively. The specimen with 50% CPW focused on the bonding vibrations within SCAACs, which were
exhibited a dense surface with presence of a small quantity of the source of compressive strength. This observation was corrobo-
non-reacted and partly reacted particles (Fig. 14a). On the other rated by the occurrence of a FTIR spectral band of Si-O-Al of greater
hand, the other specimen with 70% CPW exhibited a larger amount width. Moreover, as the CPW concentration increased from 0% to
of particles non-reacted and partly reacted (Fig. 14b). The increase 50% to 70% and the GBFS concentration declined, the band fre-
in CPW concentration from 50% to 70% was also accompanied by quency increased from 956.1 cm1 to 965.4 cm1 to 982.6 cm1,
an increased quantity of non-reacted quartz (SiO2) and poor mor- respectively. The decrease in Al levels in SCAAC matrix, and implic-
phology structure leading to a high degree of porosity [46,57,65]. itly, the reduction in strength of three-dimensional structure were
The compressive strength was adversely affected by this and the reasons for the rise in FTIR vibration frequency.
showed decreasing values of 44.8 MPa to 34.1 MPa, respectively.
It is established that an increment in CPW content and a reduction
4.9. Tensile strength
in GBFS content effect negatively on C-(A)-S-H gels formulation
and generated more partially reacted gel such as mullite and
Splitting tensile strength (TS) values of the SCAACs specimens
non-reacted particles including quartz [27].
containing various levels of CPW as binder replacement to GBFS
is enlisted in Fig. 16. At 28 days of age, the average value of three

Fig. 16. Effect of CPW replacing GBFS on tensile strength of SCAACs at age of
Fig. 15. Effect of CPW replacing GBFS on FTIR results of SCAACs. 28 days.
10 G.F. Huseien et al. / Construction and Building Materials 236 (2020) 117574

Fig. 17. Relationship between compressive strength and tensile strength.

samples was considered to determine the TS results. The TS results


revealed a lower strength values (2.9 MPa) at age of 28 days with Fig. 18. Effect of CPW replacing GBFS on FS of SCAACs at age of 28 days.
an increased content of CPW (80%) as compared to 6.4 MPa
achieved for 100% GBFS content (control sample). For other levels
of CPW replacing GBFS, it was observed that the strength values
dropped slightly, i. e., from 6.4 to 5.8, 5.3, 4.9, 4.7, 4.5, 4.2 and
3.4 MPa with the increase in CPW from 0 to 10, 20, 30, 40, 50, 60
and 70% (by mass), respectively. The reduction in CaO content with
increasing CPW content as GBFS replacement led to slow down the
rate of chemical reactions to produce C-(A)-S-H gel. Phoo-
ngernkham et al. reported [53] that the strength of alkali-
activated improved with increasing Ca content in which additional
C-S-H and C-A-S-H gels co-existed with N-A-S-H gel. This
explained the drop in TS values with increase in CPW amount as
GBFS replacement.
Fig. 17 demonstrated the relationship between compressive
strength and tensile strength of SCAACs specimens prepared with
different levels of CPW replacing GBFS. An exponential relationship
was found between compressive strength and TS with coefficient
of determination R2 value up to 0.88 as given in Eq (1). This indi-
cated a good confidence for the relationships.

TS ¼ 2:0384e0:0162P ð1Þ

R2 ¼ 0:9549 Fig. 19. WA of SCAACs prepared with various ratios of CPW to GBFS.

where

TS: tensile strength at age of 28 day, MPa


P: Compressive strength at age of 28 day, MPa

4.10. Flexural strength

Fig. 18 shows the flexural strength development of SCAACs at


age of 28 days containing different levels of CPW. The flexural
strength was influenced by the CPW levels and dropped down from
2.2 to 1.2 MPa as its contents increased from 0 to 80%, respectively.
By adding CPW in SCAACs, the flexural strength was lower than the
reference specimen (100% GBFS) for all replacement levels
between 10 and 80% and showed similar trends to strength loss
as in compressive and tensile strengths. These results were in
agreement with the previous findings [21,65] which reported that
a decrease in GBFS content as FA replacement could influence neg-
atively the formation of C-(A)-S-H product, thus reducing the
strength. The above finding of results was in agreement with the
previous studies [23,47,74] which stated that the flexural strength
value tend to drop as the GBFS content decrease in the alkali- Fig. 20. Strength loss percentage of SCAACs exposed to 10% H2SO4 solution for
activated matrix. 365 days.
G.F. Huseien et al. / Construction and Building Materials 236 (2020) 117574 11

Fig. 21. Weight loss percentage of SCAACs exposed to 10% H2SO4 solution for Fig. 22. UPV loss percentage of SCAACs exposed to 10% H2SO4 solution for 365 days,
365 days.

4.12. Sulphuric acid attack


4.11. Water absorption
Figs. 20–22 illustrated the percentage loss in strength, weight
Fig. 19 displays the water absorption (WA) results for varying and ultrasonic pulse velocity (UPV) of SCAACs exposed to 10% sul-
levels of CPW containing SCAACs. The WA of SCAACs was signifi- phuric acid solution attack, respectively. The resistance of speci-
cantly affected by the decrease in GBFS contents replacing with mens prepared with various CPW content as GBFS replacement
CWP. The WA increased from 6.8% to 10.1 and 14.1% when the (0, 10, 20, 30, 40, 50, 60, 70 and 80%) was examined. Fig. 20 pre-
CPW levels increased from 0 to 50 and 80%, respectively. The rea- sents the strength loss percentages of SCAACs immersed in 10%
sons for such findings were attributed in sections (4.5–4.8). As the H2SO4 solution for 12 months. The loss in strength decreased grad-
CPW contents increased, less dense C-(A)-S-H gel was formed ually with increasing content of CPW replacing GBFS in alkali-
which led to lower strength with high water absorption. It was activated matrix. As the CPW content increased from 0 to 80%,
reported [63] that a raise in CPW content could impact the partially the loss in strength decreased from 74% to 13.3% respectively. At
and non-reacted particles which produce more porous structure each level of CPW replacing GBFS, the strength loss was assessed.
and led to show higher water absorption in SCAACs specimens It was observed that the strength dropped to 61.4, 54.8, 46.4,
[54]. 24.9 and 15.2% as the CPW level increased from 10, 20, 50, 60 to

Fig. 23. Surface texture of SCAACs exposed to 10% H2SO4.


12 G.F. Huseien et al. / Construction and Building Materials 236 (2020) 117574

70%, respectively. Upon exposing the specimens in sulphuric acid, Acknowledgements


the presence of Ca(OH)2 in concrete reacted with SO-2 4 ions and
formed gypsum (CaSO42H2O). This product caused the expansion This research was supported and funded by UTM’s Centre of
in alkali-activated matrix which created additional cracks in the Excellence grant Q.J130000.2409.04G00. Authors would like to
interior of specimens and led to their further deterioration take the opportunity to thank the grants provider, namely Univer-
[17,75]. The loss in weight percentage of specimens exposed to sul- siti Teknologi Malaysia and Ministry of Higher Education, staff
phuric acid was also evaluated and a similar trend was observed from Materials and Structure Laboratory, School of Civil Engineer-
for all the specimens as in strength loss. The weight loss percent- ing, Universiti Teknologi Malaysia, the EM Malaysia Sdn. Bhd. and
age decreased with increasing level of CPW replacing GBFS. The EMRO Japan for their support and cooperation in conducting this
increase in CPW level from 0% to 10, 20, 30, 40, 50, 60, 70 and research.
80% led to drop the weight loss percentage from 2.2 to 1.7, 1.2,
0.98, 0.76, 0.44 and 0.39, respectively (Fig. 21). The internal deteri-
oration of specimens immersed for one year in acid solution was Appendix A. Supplementary data
also evaluated by using UPV test. Fig. 22 shows the loss in UPV per-
centage values. It was observed that the percentage dropped from Supplementary data to this article can be found online at
20.2% to 15.3, 13.4, 9.8, 5.6, 3.3 and 3.1% with increase in WCP from https://doi.org/10.1016/j.conbuildmat.2019.117574.
0% to 10, 20, 30, 40, 50, 60, 70 and 80%, respectively. The decrease
in surface deterioration and cracks with the increase in CPW con-
References
tents can be clearly seen in Fig. 23. The high calcium content in
control sample (100% GBFS) compared to other matrixes resulted [1] G. Long, Y. Gao, Y. Xie, Designing more sustainable and greener self-
in higher gypsum content. Therefore, the higher resistance to sul- compacting concrete, Constr. Build. Mater. 84 (2015) 301–306.
phuric acid attack and loss less than 13.3% of compressive strength [2] P. Domone, Self-compacting concrete: an analysis of 11 years of case studies,
Cem. Concr. Compos. 28 (2) (2006) 197–208.
was observed in specimens containing high CPW content. Allah- [3] G.F. Huseien, J. Mirza, M. Ismail, Effects of high volume ceramic binders on
verdi and Skvara [76] reported the sodium silicate activated spec- flexural strength of self-compacting geopolymer concrete, Adv. Sci. Lett. 24 (6)
imens containing 50% GBFS immersed in acid solution and (2018) 4097–4101.
[4] H. Okamura, Self-compacting high-performance concrete, Concr. Int. 19 (7)
indicated that the free Ca++ and Na+ in the formed alkali- (1997) 50–54.
activated network were negative to acid resistance. It was demon- [5] C. Shi et al., A review on mixture design methods for self-compacting concrete,
strated that ‘‘the exchange reaction of Ca++ and Na+ with OH ions Constr. Build. Mater. 84 (2015) 387–398.
[6] S. Subasßı, H. Öztürk, M. Emiroğlu, Utilizing of waste ceramic powders as filler
results in their loss, thus further degrading the gels”. In addition,
material in self-consolidating concrete, Constr. Build. Mater. 149 (2017) 567–
de-alumination could occur due to acid attack prone on Si-O-Al 574.
bonds, causing the composition changes and structure of ASs net- [7] G.F. Huseien et al., Geopolymer mortars as sustainable repair material: a
work [77]. comprehensive review, Renew. Sustain. Energy Rev. 80 (2017) 54–74.
[8] I. García-Lodeiro et al., Compatibility studies between NASH and CASH gels.
Study in the ternary diagram Na2O–CaO–Al2O3–SiO2–H2O, Cem. Concr. Res.
41 (9) (2011) 923–931.
5. Conclusions [9] H. Abdel-Gawwad et al., Recycling of concrete waste to produce ready-mix
alkali activated cement, Ceram. Int. 44 (6) (2018) 7300–7304.
1. All SCAAC specimens prepared with 50% and higher CPW in this [10] G.F. Huseien, K.W. Shah, A.R.M. Sam, Sustainability of nanomaterials based
self-healing concrete: an all-inclusive insight, J. Build. Eng. (2019).
study achieved adequate flow, passing ability and resistance to [11] G.F. Huseien et al., Synthesis and characterization of self-healing mortar with
segregation. These results indicated that the flow and passing modified strength, J. Teknol. 76 (1) (2015).
ability of SCAAC increased with the increment of adding CPW [12] I.O. Hassan et al., Flow characteristics of ternary blended self-consolidating
cement mortars incorporating palm oil fuel ash and pulverised burnt clay,
in the matrix. Contrast to these results, the SCAACs’ resistance Constr. Build. Mater. 64 (2014) 253–260.
to segregation decreased with increasing CPW contents. How- [13] M.A. Asaad et al., Improved corrosion resistance of mild steel against acid
ever, incorporating GBFS and CPW led to achieve the aims of activation: impact of novel Elaeis guineensis and silver nanoparticles, J. Ind.
Eng. Chem. 63 (2018) 139–148.
this study and produced high performance SCAAC. [14] G.F. Huseien et al., The effect of sodium hydroxide molarity and other
2. Inclusion of CPW enhanced both the initial and final setting parameters on water absorption of geopolymer mortars, Indian J. Sci. Technol.
times of SCAAC samples. 9 (48) (2016) 1–7.
[15] C. Ouellet-Plamondon, G. Habert, Life cycle assessment (LCA) of alkali-
3. Incorporating CPW as raw material in SCAACs showed excellent activated cements and concretes, in: Handbook of Alkali-Activated Cements,
performance in hardened state. All SCAAC specimens presented Mortars and Concretes, Elsevier, 2015, pp. 663–686.
acceptable compressive strength between 30 and 50 MPa. How- [16] B.C. McLellan et al., Costs and carbon emissions for geopolymer pastes in
comparison to ordinary portland cement, J. Cleaner Prod. 19 (9–10) (2011)
ever, the strength dropped as the CPW contents increased in
1080–1090.
SCAAC matrix. Similar trends were also found with FS and TS. [17] G.F. Huseien et al., Evaluation of alkali-activated mortars containing high
4. According to XRD, SEM and FTIR microstructure results, the for- volume waste ceramic powder and fly ash replacing GBFS, Constr. Build.
mation of C-S-H and C-A-S-H gels was influenced by high con- Mater. 210 (2019) 78–92.
[18] D. Hardjito et al., Factors influencing the compressive strength of fly ash-based
tents of SiO2 in CPW. The gels formation decreased as the CPW geopolymer concrete, Civil Eng. Dimens. 6 (2) (2004) 88–93.
contents increased which resulted in poor structure, low [19] M. Soutsos et al., Factors influencing the compressive strength of fly ash based
strength and high porosity. geopolymers, Constr. Build. Mater. 110 (2016) 355–368.
[20] G.F. Huseien, M. Ismail, J. Mirza, Influence of curing methods and sodium
5. Inclusion of CPW in alkali-activated matrix as GBFS replace- silicate content on compressive strength and microstructure of multi blend
ment led to enhance the durability of SCCACs exposed to sul- geopolymer mortars, Adv. Sci. Lett. 24 (6) (2018) 4218–4222.
phuric acid attack. [21] G.F. Huseiena et al., Performance of sustainable alkali activated mortars
containing solid waste ceramic powder, Chem. Eng. (2018) 63.
[22] P. Duan et al., Development of fly ash and iron ore tailing based porous
geopolymer for removal of Cu (II) from wastewater, Ceram. Int. 42 (12) (2016)
13507–13518.
Declaration of Competing Interest [23] G.F. Huseien et al., Synergism between palm oil fuel ash and slag: Production
of environmental-friendly alkali activated mortars with enhanced properties,
The authors declare that they have no known competing finan- Constr. Build. Mater. 170 (2018) 235–244.
[24] G.F. Huseien et al., Utilizing spend garnets as sand replacement in alkali-
cial interests or personal relationships that could have appeared activated mortars containing fly ash and GBFS, Constr. Build. Mater. 225
to influence the work reported in this paper. (2019) 132–145.
G.F. Huseien et al. / Construction and Building Materials 236 (2020) 117574 13

[25] M.O. Yusuf et al., Evolution of alkaline activated ground blast furnace slag– [52] M.B. Karakoç et al., Mechanical properties and setting time of ferrochrome slag
ultrafine palm oil fuel ash based concrete, Mater. Des. 55 (2014) 387–393. based geopolymer paste and mortar, Constr. Build. Mater. 72 (2014) 283–292.
[26] A.M. Rashad, Properties of alkali-activated fly ash concrete blended with slag, [53] T. Phoo-ngernkham et al., High calcium fly ash geopolymer mortar containing
Iran. J. Mater. Sci. Eng. 10 (1) (2013) 57–64. Portland cement for use as repair material, Constr. Build. Mater. 98 (2015)
[27] G.F. Huseiena et al., Effect of binder to fine aggregate content on performance 482–488.
of sustainable alkali activated mortars incorporating solid waste materials, [54] Z. Kubba et al., Impact of curing temperatures and alkaline activators on
Chem. Eng. (2018) 63. compressive strength and porosity of ternary blended geopolymer mortars,
[28] S.V. Patankar, Y.M. Ghugal, S.S. Jamkar, Effect of concentration of sodium Case Stud. Constr. Mater. 9 (2018) e00205.
hydroxide and degree of heat curing on fly ash-based geopolymer mortar, [55] T. Phoo-ngernkham et al., Effects of sodium hydroxide and sodium silicate
Indian J. Mater. Sci. 2014 (2014). solutions on compressive and shear bond strengths of FA–GBFS geopolymer,
[29] D. Khale, R. Chaudhary, Mechanism of geopolymerization and factors Constr. Build. Mater. 91 (2015) 1–8.
influencing its development: a review, J. Mater. Sci. 42 (3) (2007) 729–746. [56] W.D. Rickard et al., Assessing the suitability of three Australian fly ashes as an
[30] N. Ranjbar et al., Compressive strength and microstructural analysis of fly aluminosilicate source for geopolymers in high temperature applications,
ash/palm oil fuel ash based geopolymer mortar under elevated temperatures, Mater. Sci. Eng., A 528 (9) (2011) 3390–3397.
Constr. Build. Mater. 65 (2014) 114–121. [57] G.F. Huseien et al., Properties of ceramic tile waste based alkali-activated mortars
[31] M.A. Salih et al., Development of high strength alkali activated binder using incorporating GBFS and fly ash, Constr. Build. Mater. 214 (2019) 355–368.
palm oil fuel ash and GGBS at ambient temperature, Constr. Build. Mater. 93 [58] G.F. Huseien et al., Influence of different curing temperatures and alkali
(2015) 289–300. activators on properties of GBFS geopolymer mortars containing fly ash and
[32] F. Pacheco-Torgal, S. Jalali, Reusing ceramic wastes in concrete, Constr. Build. palm-oil fuel ash, Constr. Build. Mater. 125 (2016) 1229–1240.
Mater. 24 (5) (2010) 832–838. [59] W. Zhou et al., A comparative study of high-and low-Al2O3 fly ash based-
[33] B. Huang, Q. Dong, E.G. Burdette, Laboratory evaluation of incorporating waste geopolymers: the role of mix proportion factors and curing temperature,
ceramic materials into Portland cement and asphaltic concrete, Constr. Build. Mater. Des. 95 (2016) 63–74.
Mater. 23 (12) (2009) 3451–3456. [60] M. Sitarz, W. Mozgawa, M. Handke, Vibrational spectra of complex ring silicate
[34] F. Pacheco-Torgal, S. Jalali, Compressive strength and durability properties of anions—method of recognition, J. Mol. Struct. 404 (1–2) (1997) 193–197.
ceramic wastes based concrete, Mater. Struct. 44 (1) (2011) 155–167. [61] J.W. Phair, J. Van Deventer, J. Smith, Mechanism of polysialation in the
[35] Z. Abdollahnejad et al., Mix design, properties and cost analysis of fly ash- incorporation of zirconia into fly ash-based geopolymers, Ind. Eng. Chem. Res.
based geopolymer foam, Constr. Build. Mater. 80 (2015) 18–30. 39 (8) (2000) 2925–2934.
[36] P. Nath, P.K. Sarker, Effect of GGBFS on setting, workability and early strength [62] X. Guo, H. Shi, W.A. Dick, Compressive strength and microstructural
properties of fly ash geopolymer concrete cured in ambient condition, Constr. characteristics of class C fly ash geopolymer, Cem. Concr. Compos. 32 (2)
Build. Mater. 66 (2014) 163–171. (2010) 142–147.
[37] G.F. Huseien et al., Effect of metakaolin replaced granulated blast furnace slag [63] S. Efnarc, Guidelines for Self-compacting Concrete, Association House, London,
on fresh and early strength properties of geopolymer mortar, Ain Shams Eng. J. UK, 2002, p. 34.
(2016). [64] ASTM, C., Standard Test Methods for Chemical Resistance of Mortars, Grouts,
[38] N.H. Roslan et al., Performance of steel slag and steel sludge in concrete, and Monolithic Surfacing and Polymer Concretes. 2012.
Constr. Build. Mater. 104 (2016) 16–24. [65] G.F. Huseien et al., Waste ceramic powder incorporated alkali activated
[39] A. Çevik et al., Effect of nano-silica on the chemical durability and mechanical mortars exposed to elevated temperatures: performance evaluation, Constr.
performance of fly ash based geopolymer concrete, Ceram. Int. 44 (11) (2018) Build. Mater. 187 (2018) 307–317.
12253–12264. [66] T. Sugama, L. Brothers, T. Van de Putte, Acid-resistant cements for geothermal
[40] Baraldi, L., M.-M.E.S. by ACIMAC, World production and consumption of wells: sodium silicate activated slag/fly ash blends, Adv. Cem. Res. 17 (2)
ceramic tiles. America, 2016. 1(9.2): p. 7.7. (2005) 65–75.
[41] A.A. Hussein et al., Performance of nanoceramic powder on the chemical and [67] H. Alanazi, J. Hu, Y.-R. Kim, Effect of slag, silica fume, and metakaolin on
physical properties of bitumen, Constr. Build. Mater. 156 (2017) 496–505. properties and performance of alkali-activated fly ash cured at ambient
[42] R. Senthamarai, P.D. Manoharan, D. Gobinath, Concrete made from ceramic temperature, Constr. Build. Mater. 197 (2019) 747–756.
industry waste: durability properties, Constr. Build. Mater. 25 (5) (2011) [68] G.F. Huseien et al., Effects of POFA replaced with FA on durability properties of
2413–2419. GBFS included alkali activated mortars, Constr. Build. Mater. 175 (2018) 174–
[43] M. Samadi et al., Properties of mortar containing ceramic powder waste as 186.
cement replacement, J. Teknol. 77 (12) (2015) 93–97. [69] S. Kumar, R. Kumar, S. Mehrotra, Influence of granulated blast furnace slag on
[44] N.H.A.S. Lim et al., Microstructure and strength properties of mortar the reaction, structure and properties of fly ash based geopolymer, J. Mater.
containing waste ceramic nanoparticles, Arab. J. Sci. Eng. 43 (10) (2018) Sci. 45 (3) (2010) 607–615.
5305–5313. [70] F. Puertas et al., Alkali-activated fly ash/slag cements: strength behaviour and
[45] M. Mohit, Y. Sharifi, Thermal and microstructure properties of cement mortar hydration products, Cem. Concr. Res. 30 (10) (2000) 1625–1632.
containing ceramic waste powder as alternative cementitious materials, [71] W. Mozgawa, J. Deja, Spectroscopic studies of alkaline activated slag
Constr. Build. Mater. 223 (2019) 643–656. geopolymers, J. Mol. Struct. 924 (2009) 434–441.
[46] H. Mohammadhosseini et al., Enhanced performance of green mortar [72] D. Ravikumar, S. Peethamparan, N. Neithalath, Structure and strength of NaOH
comprising high volume of ceramic waste in aggressive environments, activated concretes containing fly ash or GGBFS as the sole binder, Cem. Concr.
Constr. Build. Mater. 212 (2019) 607–617. Compos. 32 (6) (2010) 399–410.
[47] M.H. Al-Majidi et al., Development of geopolymer mortar under ambient [73] I. Ismail et al., Modification of phase evolution in alkali-activated blast furnace
temperature for in situ applications, Constr. Build. Mater. 120 (2016) 198–211. slag by the incorporation of fly ash, Cem. Concr. Compos. 45 (2014) 125–135.
[48] N. Lee, E. Kim, H.-K. Lee, Mechanical properties and setting characteristics of [74] S. Song, H.M. Jennings, Pore solution chemistry of alkali-activated ground
geopolymer mortar using styrene-butadiene (SB) latex, Constr. Build. Mater. granulated blast-furnace slag, Cem. Concr. Res. 29 (2) (1999) 159–170.
113 (2016) 264–272. [75] W. Zhang et al., The degradation mechanisms of alkali-activated fly ash/slag
[49] M.O. Yusuf, Performance of slag blended alkaline activated palm oil fuel ash blend cements exposed to sulphuric acid, Constr. Build. Mater. 186 (2018)
mortar in sulfate environments, Constr. Build. Mater. 98 (2015) 417–424. 1177–1187.
[50] M.O. Yusuf et al., Evaluation of slag-blended alkaline-activated palm oil fuel [76] A. Allahverdi, F. Skvara, Sulfuric acid attack on hardened paste of geopolymer
ash mortar exposed to the sulfuric acid environment, J. Mater. Civ. Eng. 27 (12) cements-Part 1. Mechanism of corrosion at relatively high concentrations,
(2015) 04015058. Ceram. Silik. 49 (4) (2005) 225.
[51] G.F. Huseien et al., Compressive strength and microstructure of assorted [77] L. Gu, T. Bennett, P. Visintin, Sulphuric acid exposure of conventional concrete
wastes incorporated geopolymer mortars: effect of solution molarity, and alkali-activated concrete: assessment of test methodologies, Constr. Build.
Alexandria Eng. J. 57 (4) (2018) 3375–3386. Mater. 197 (2019) 681–692.

You might also like