Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Cement and Concrete Composites 147 (2024) 105417

Contents lists available at ScienceDirect

Cement and Concrete Composites


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

Alkali thermal fusion: A prospective route to enhance the reactivity of


low-grade clay and utilize as supplementary cementitious material (SCM)
Ishrat Baki Borno , Nithya Nair , Warda Ashraf *
Department of Civil Engineering, The University of Texas at Arlington, Nedderman Hall, 416 Yates St., Arlington, TX 76019, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Common clay deposits contain a variety of minerals that provide low and unpredictable reactivity even after
Pozzolan calcination. This study aims to enhance the reactivity of such low-grade clay via alkali thermal fusion, where clay
Amorphous was co-calcined with different dosages of NaOH to achieve superior reactivity. The effects of such co-calcination
X-ray diffraction
on the formation of reactive aluminosilicates were monitored using high-temperature in-situ X-ray Diffraction
Dehydroxylation
Dissolution
(XRD). Additionally, the role of different alkali dosages on the characteristics of the calcined clays was studied
Co-calcination using ex-situ XRD, FTIR, ICP-OES, and 29Si NMR. The co-calcination process significantly enhanced the Al and Si
dissolution from the calcined clays, which subsequently improved their pozzolanic properties. All of the alkali-
fused SCMs satisfied the modified strength activity index (SAI) requirement and the lower dosages of alkali-
containing samples (5 % NaOH) showed 50 % reduction in expansion due to alkali silica reaction.

1. Introduction to be the kaolinite content [19]. Based on the kaolinite content, common
clay deposits can be categorized as high (>65 % kaolinite content),
Concrete produced using Ordinary Portland Cement (OPC) as the medium (40 % ~ 65 % kaolinite content), and low grade (<40 %
primary binder has significant negative impact on the environment [1, kaolinite content) kaolin clays [20–23]. The high-grade kaolin sources
2]. The negative environmental impacts of concrete can be reduced by are fewer in number and not equally distributed all over the world [24],
partially replacing OPC with supplementary cementitious materials compared to its application in various fields for instance the cement,
(SCMs) [2]. Studies have shown that SCMs can be used to improve paper, whiteware, and refractories industries which are reflected in
concrete’s workability, long-term strength, durability, and sustainability limited availability and relatively high prices for metakaolin in com­
[3–6]. The availability of natural pozzolans such as tuffs and zeolites are parison with other SCMs [21]. Considering the availability, trans­
confined to certain geographical locations. The conventional SCMs, portation cost, and energy-intensive clay purification expenses, locally
including coal fly/bottom ash, different types of slags, solid waste available low-grade kaolin clays can be a suitable source for SCM pro­
incineration ash, ashes from agricultural sources, and wollastonite are duction [20].
not readily available throughout the world [7–11]. Additionally, the The aim of this study is to utilize a locally available low-grade clay as
availability of one of the most widely used SCM-coal fly ash, is currently a supplementary cementitious material (SCM) by adopting the alkali
declining due to the closing of coal-based power plants. Calcined clay fusion route. In the alkali fusion process (also often addressed as alkali
can be a viable candidate in this regard because of its wide availability thermal fusion in literature), the aluminosilicate precursors are co-
across the world [12–14]. The common clay deposits contain a blend of calcined in presence of Alkali or Alkaline Earth Metal (AAEM) hydrox­
primarily kaolinite, montmorillonite, and illite/smectite type clay ide. Alkali fusion is traditionally used for extracting metals from solid
minerals in addition to various non-clay or associated minerals (e.g., wastes [25–29]. This process involves solid-state reactions between
quartz, feldspar, etc.). Among the clay minerals, kaolinite shows supe­ alumina or silica-bearing minerals with hydroxides or carbonates of
rior reactivity after thermal, mechanical or chemical activation [15–17]. AAEM species at a temperature ranging from 180 to 950 ◦ C [30–32].
The most reactive calcined clay product is metakaolin, formed by the During this process, the AAEM ions decompose and subsequently modify
de-hydroxylation of kaolinite at or above 600 ◦ C [15,16,18]. The the aluminosilicate or silicate network to increase the degree of
limiting factor in the pozzolanic reactivity of calcined clay is considered amorphization [33] and form new phases with higher dissolution rates

* Corresponding author.
E-mail address: warda.ashraf@uta.edu (W. Ashraf).

https://doi.org/10.1016/j.cemconcomp.2023.105417
Received 31 July 2023; Received in revised form 4 December 2023; Accepted 24 December 2023
Available online 29 December 2023
0958-9465/© 2024 Elsevier Ltd. All rights reserved.
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

[34]. The AAEM species can play a dual role in the aluminosilicate sample according to Table 2 and mixed for 90 min in the planetary ball
network – as charge compensators to balance the negative charge caused mill at a speed of 2520 rpm for uniform blending and then calcined at
by the substitution of Si by Al, or as network modifiers where they 600 ◦ C for 60 min in a muffle furnace. The ball mill contained four nylon
disrupt Si–O–Si bonds and form non-bridging oxygens (NBOs) [35,36]. jars with 8 pieces of zirconia balls of 4 different sizes per jar. The clays
Alkali fusion was also used in the past to produce zeolites from were ground again in the ball mill for around 150 min to ensure uniform
relatively poor-quality coal fly ash [37], clay [32,38,39], and various particle size. For this laser particle size analyzer was used. In this study,
types of incineration ash [40]. Previous studies have shown that laboratory-grade isopropanol (2 % by weight) was used as a grinding aid
co-calcination with AAEM additives led to the conversion of various based on the findings of previous studies [43,44]. The obtained powder
crystalline minerals, including gehlenite, mullite, quartz, and talc into was used as the SCM in this study. The calcination temperature was
soluble and amorphous aluminosilicates [34,37,41]. Accordingly, the selected based on our previous work that showed nepheline formation at
reactivity of low-grade clay can potentially be improved by such an 750 ◦ C [43], and the literature data where low-reactive granite was
alkali-fusion or co-calcination process. A recent study [42] evaluated the utilized via alkali-fusion at 650 ◦ C [45].
effectiveness of co-calcining clays with dolomites to produce SCM. The
authors concluded that the co-calcination of kaolinitic clays with dolo­
mites enables applying relatively high temperatures (~950 ◦ C) while 2.3. Experimental methods
also achieving enhanced reactivity. In our recent article [43], we
showed that co-calcination (alkali fusion) with a high dosage of NaOH 2.3.1. Density measurement
(more than 10 %) and at a relatively high temperature negatively affects Density of the raw and alkali fused SCM was measured using the bulk
the reactivity of kaolinite-rich clays due to the alternation of Al-species. density method where powder samples were taken, compressed with 10
The past study also revealed that alkali fusion can potentially reduce the MPa pressure with a pellet press to remove air void, and then the mass
dehydroxylation temperature requirement of 2:1 clay mineral (e.g., and the dimension were recorded. From these data, density was calcu­
bentonite) [43]. Based on those previous findings, in this present study, lated. Each data is an average of two consecutive readings.
the effects of alkali fusion on low-grade clay at a relatively low tem­
perature (~600 ◦ C) are evaluated. 2.3.2. In-situ high temperature X-ray diffraction (HT-XRD)
With the primary goal of evaluating the effectiveness of alkali fusion In-situ high-temperature X-ray diffraction patterns were collected
to produce SCM from low-grade clay, the specific objectives of this using a Rigaku SmartLab II. Data were collected for the 2θ range from 5◦
article are as follows: (i) to understand how the co-calcination with to 60◦ with a step size of 0.03◦ (2θ) per second. The temperature was
sodium hydroxides alters the aluminosilicate (or silicate) network of increased from room temperature to 1000 ◦ C at a rate of 275.15 K/min.
impure clay, (ii) to investigate the effects of co-calcination on Al and Si The HT-XRD patterns were collected at 15-min time intervals. To obtain
dissolution from calcined clays, and (iii) to evaluate the performance of uniform powder samples for this measurement, the raw clay samples
low-grade clays calcined with and without sodium hydroxides as SCM. were dried at 110 ◦ C for 24 h, then ground with and without NaOH in a
These objectives were achieved by using a wide range of experimental planetary ball mill for 90 min. Total 32 XRD patterns were collected for
techniques, including Fourier transform infrared spectroscopy (FTIR), X- each sample. The background was identified for the first XRD pattern at
ray diffraction (XRD), nuclear magnetic resonance (NMR), and ther­ room temperature and the same background was deducted from each of
mogravimetric analysis (TGA). The impacts of alkali-fused clays on the the spectra. The mineral phases were identified using the first and last
compressive strengths and Alkali-Silica Reactivity (ASR) were also XRD patterns using the software Match!
evaluated.
2.3.3. X-ray diffraction
2. Materials and methods X-ray Diffraction (XRD) patterns of the powdered samples were ob­
tained using a Cu-Kα source (Bruker D8 ADVANCE). The diffraction
2.1. Raw materials patterns were obtained for the 2θ range from 5◦ to 60◦ using a step size
of 0.03 (2θ) per second. To observe the effect of the alkali fusion on the
The low-grade clay was collected from local area in Irving, Texas, aluminosilicate network, the samples prepared in section 2.2 was
USA. The chemical composition of the clay was determined using X-ray directly utilized without further modification. For determining the
fluorescence (XRF) and is presented in Table 1. This clay was found to mineralogy of the hydration product, the 28 days-cured (curing condi­
contain 22 % kaolinite, 19 % illite, 57 % quartz, a trace amount of tion described in section 2.3.8) paste samples were crushed manually
calcite, and other impurities, as measured using TGA and XRD (Rietveld (using mortar and pestle) and then sieved through a number 200 sieve
refinement was performed using 10 % corundum as the internal (0.074 mm opening) and the obtained powder was used for XRD. The
standard).
Table 2
Material mix details.
2.2. Sample preparation
Sample Id Calcination condition Clay (wt %) NaOH added (wt %)
The moisture content of the raw clay was found to be 15 % when Raw Not calcined 100 0
calculated using the following formula. Control Calcined 0
5 % NaOH 5
wt of wet clay-wt of dry clay 10 % NaOH 10
Moisture (%) = × 100 15 % NaOH 15
wt of dry clay
20 % NaOH 20
Different NaOH (solid pellet, >98 % pure, purchased from VWR and 25 % NaOH 25
used without further alteration) dosages were added to the raw clay

Table 1
Oxide contents of the raw materials (mass %), LOI: Loss on Ignition.
Oxides (mass %) SiO2 Al2O3 Fe2O3 CaO K2O MgO TiO2 SO3 Na2O P2O5 LOI

Low-grade clay 58.7 21.6 8.18 3.74 2.95 1.85 1.2 0.66 0.63 0.23 10.5
OPC 21.2 3.87 3.1 65.3 0.68 0.82 0.137 4.28 – 0.61 3.43

2
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

analysis was carried out using MATCH! Software with Powder Diffrac­ 2.3.9. Thermogravimetric analysis (TGA)
tion File (PDF) from International Centre for Diffraction Data (ICDD) The commercially available TGA 550 TA instrument was used for this
database. purpose. In this study, paste samples were first ground using a mortar
and pestle, and then approximately 35–40 mg of ground powder sample
2.3.4. FTIR was loaded into a platinum pan/crucible and kept in an isothermal
Fourier Transform Infrared Spectroscopy (FTIR) was carried out condition at around 25 ◦ C for 5 min. Then the temperature of the TGA
using a Nicolet iS50 FTIR equipped with an attenuated total reflectance chamber was increased to 980 ◦ C with a ramp of 15 ◦ C per minute. N2
(ATR) accessory. The frequency range was 400–4000 cm− 1 at a resolu­ gas was purged during the entire process. For determining portlandite
tion of 4 cm− 1. Each spectrum presented in this paper was an average of consumption, the mass loss over 400 ◦ C–500 ◦ C temperature range was
32 scans. used as per the methods published elsewhere [11,47]. For a few initial
batches, three replicate samples were tested through TGA to validate for
2.3.5. 29Si Solid state nuclear magnetic resonance (NMR) any deviation in carbonation across samples. The test result deviations
All the 29Si solid-state NMR experiments were conducted at the SCS were less than 2 % by weight of the samples. Due to the low deviation,
NMR Facility of the University of Illinois at Urbana-Champaign, at room TGA was performed with only one sample for the remainder of the
temperature, at 7.05 T on a Varian Unity Inova 300 MHz spectrometer samples.
operating at a resonance frequency of ν0 (29Si) = 59.6 MHz. A Varian/ The portlandite consumptions by the calcined clays were utilized for
Chemagnetics 7.5 mm double-resonance APEX HX magic-angle spinning depicting the pozzolanic reactivity [48–50]. The samples obtained from
(MAS) probe was used for all the MAS experiments under a spinning rate the R3 test after 7 days were utilized to determine the portlandite con­
of 4 kHz and TPPM 1H decoupling. All samples were finely ground and sumption. The hydration reaction of the samples was arrested following
packed into 4 mm o. d. standard zirconia rotors (typically around the solvent exchange method and then the thermogravimetric analysis
210–580 mg). Experimental silicon chemical shift referencing, pulse was performed to determine the portlandite consumption of the samples
calibration and setup were done using powdered octakis (trimethylsi­ [51]. Initially, 66.67 g of calcium hydroxide [Ca(OH)2] per 100 g of total
loxy)silsesquioxane (Q8M8), which has a chemical shift of 11.45 ppm, solid from the R3 test was added to the mixture as per the R3 test stan­
relative to the primary standard, TMS at 0 ppm. The 29Si pulse width dard requirement.
used was 1.5 μs, corresponding to a 45-degree pulse. A recycle delay of
30s was used and 1920 scans were acquired for each sample. 2.3.10. Compressive strength
The compressive strength was measured for 50 mm × 50 mm × 50
2.3.6. ICP-OES mm mortar cube prepared as per ASTM C311 [52] where 20 % by wt of
The dissolution rates for Al and Si were determined using inductively Ordinary Portland Cement (OPC) Type I/II was replaced by SCM
coupled plasma-optical emission spectrometry (ICP-OES). For this (alkali-fused calcined clays). ASTM-graded standard sand and deionized
measurement, 0.1 g of the solid precursor (alkali-fused powder prepared water were used to prepare mortar cubes. The water to binder ratio was
as per section 2.1) was mixed with water at a liquid: solid ratio of 400:1. 0.485. Samples were cured in saturated lime water at room temperature
The pH of the water was adjusted to 13.5 using NaOH. The solutions (23 ◦ C ± 1 ◦ C). Compressive strengths were determined after 7, 14, 28,
were prepared in HDPE centrifuge tubes and placed on a tube rocker (at and 56 days of curing as per ASTM C109 [53].
25 rpm) that continuously agitated the samples. Each sample mixtures Strength Activity Index (SAI) is a widely used method to evaluate the
were filtered at specific durations and the solution was used for ICP-OES reactivity of supplementary cementitious materials (SCMs), determined
measurements. A total of six standard solutions (0 ppm, 0.5 ppm, 5 ppm, as per ASTM C618. In this study, the water-to-binder was kept constant,
10 ppm, 20 ppm, and 50 ppm) were used for calibrating ICP-OES and hence it is referred as ‘modified’ SAI [54].
measurements. The selected wavelengths for Al and Si were 394.403
nm and 288.158 nm, respectively. Each result presented in this study is 2.3.11. Heat of hydration
an average of three measurements. The filtered solids were dried in an Heat releases of the paste samples caused by hydration of the paste
oven at 40 ◦ C for 24 h and evaluated using XRF and XRD. samples of pure OPC with and without alkali fused SCMs (20 %
replacement of OPC by wt) were monitored using an isothermal calo­
2.3.7. X-ray fluorescence (XRF) rimeter (TAM Air, TA instruments). An in-situ setup was used. The
XRF analysis of the powder samples were performed using a Rigaku water-to-binder ratio was 0.485. The heat release data was continuously
NEX CG spectrometer. According to the instrument manual, 4 g of taken for 168 h. This experiment was carried out at 25 ◦ C.
sample is recommended to use per analysis. The powder samples were
pulverized with a mortar pestle and passed through a #200 sieve. 4 g of 2.3.12. Alkali-silica reaction (ASR)
this fine powder was then pressed into pellets using the pellet press. The alkali silica reaction (ASR) test was done as per the guidelines of
NIST standard reference material was used for calibrating the XRF. ASTM C1567 [55]. Type 33 alkali-borosilicate glass, supplied by NBS,
Vitro minerals was used as the reactive aggregates. The sample prepa­
2.3.8. R3 test ration and curing regime were done as per the ASTM standard recom­
The R3 test method was carried out as per standard ASTM C1897 mendations. The expansion due to ASR was measured in terms of change
[46]. Each of the calcined clay was mixed with Ca(OH)2 and CaCO3 in length by an average of three specimens for each binder-aggregate
(clay: Ca(OH)2 = 1:3 and CaCO3: clay = 1: 2). The pore solution was combination daily for 16 days.
prepared by dissolving 4.00 g of KOH and 20.0 g of K2SO4 in 1.00 l of
deionized water. The solution was added to the solid mixture where the 2.3.13. Scanning electron microscopy (SEM) and energy dispersive X-ray
solution to solid ratio was 1.2. All the materials along with vials were spectroscopy (EDS)
kept at 40 ◦ C oven at least for 24 h before mixing. After mixing, 15 g ± For Backscattered Electrons (BSE) imaging, the alkali fused pre­
0.01 g sample was transferred to the vials and placed in the isothermal cursors (powder) and 28 day paste samples (after arresting the hydration
calorimeter (TAM AIR) at 40 ◦ C. For this measurement, deionized water and drying in a vaccum dessicator) were impregnated in an epoxy resin
was used as a reference. The measurements were collected for 168hr. and then lapped/polished to obtain a mirror-like surface finish. The
The isothermal calorimeter measurements were repeated twice for the polished samples were coated with Gold (Au)-Platinum (Pt) before
first three samples. The standard deviation between the measurements capturing the SEM images. The BSE images were obtained using the
was found to be less than 1 %. Therefore, for the remaining measure­ Hitachi 3000 N SEM. The instrument was operated in high-vacuum
ments, only one sample was evaluated. mode with a 25-kV accelerated voltage and a working distance of

3
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

about 15 mm. The EDS data points were collected using the same Table 3
working distance. Particle size and density of the samples.
Sample d10 d50 d90 Bulk density (g/ Std. dev. (g/
3. Results name (μm) (μm) (μm) cc) cc)

Raw 1.15 11.08 40.42 1.02 0.03


3.1. Physical properties Control 1.12 13.61 41.09 1.07 0.015
5 % NaOH 1.16 9.7 35.56 1.08 0.02
10 % NaOH 1.41 13.61 41.98 1.14 0.025
The BSE images of the alkali-fused precursors (control, 5 %, and 25 %
15 % NaOH 1.62 15.69 37.47 1.16 0.025
NaOH) are presented with EDS points Fig. 1a-j. All of the samples con­ 20 % NaOH 1.14 11.08 33.36 1.14 0.02
tained unaltered quartz (SiO2) (Fig. 1c,f,j). In Fig. 1a, there were two 25 % NaOH 0.5 7.33 30.13 1.17 0.03
distinct layers of clay particles, where the inner particles are denser and
more compact compared to the outer layer. This could be the result of
grinding after calcination. The outer layer particles got affected and decomposition temperature observed here compared to the literature
produced finer and more dispersed particles while some of the inner [16] is due to the fast-heating rate used in the experiments. However,
layer particles remained as it is. The 5 % NaOH (Figs. 1e) and 25 % the dehydroxylation temperature range could also vary within a range
NaOH (Fig. 1h) alkali fused sample had Na bound with the clay particles. from 550 to 900 ◦ C, depending on different types of illite [56,57].
In 25 % NaOH containing sample, nepheline formation was identified Mullite peaks appeared due to the recrystallization of kaolin at around
(Fig. 1i), where the Na to Al ratio was higher than the alkali fused clay 850 ◦ C. Mullite can form at around 800–1400 ◦ C as a result of solid-state
particles (Na/Al was around 2.28 for nepheline whereas Na/Al ranged reactions between silica and alumina based on the mutual diffusion of
from 0.26 to 0.66 for 25 % NaOH fused clay). Particle size distributions Al3+, Si4+, and O2− ions among particles [58]. For the sample containing
of the materials were determined through laser diffraction. Fig. 1k 10 % NaOH (Fig. 2b), kaolinite, and illite peaks were diminished at
shows the differential particle size distribution of the alkali-fused sam­ around 480 ◦ C and 620 ◦ C, respectively. The formation of mullite was
ples. The alkali fused batches with NaOH dosage of higher than 15 % not observed to form in this sample. However, nepheline (crystalline
had relatively smaller diameter (d50) particles (Table 3), indicating that sodium aluminosilicate, Na3K(Al4Si4O16)) peaks started appearing at
the alkali fusion affected the hardness of the particles. While the overall around 850 ◦ C, which is more than 100 ◦ C after the dehydroxylation of
particle size distribution followed similar pattern, it was not possible to illite. In the later part of this study (section 3.3.1), the crystallization
achieve the same d50 after 150 min of ball milling. temperature of nepheline was found to be dependent on the NaOH
dosage. Regardless, based on these findings, the addition of NaOH can
be a potential pathway to reduce the dehydroxylation temperature of
3.2. Effects of NaOH on the calcination process clay minerals.
The DTG plots of the common clays with various NaOH dosages are
Fig. 2 shows the in-situ XRD patterns of the clay collected while given in Fig. 3. The dehydration peak of the NaOH-containing samples
heating it up to 1000 ◦ C. The decomposition and formation of different showed relatively higher weight loss indicating the presence of higher
phases were marked using red and black arrows, respectively. Without moisture compared to the control sample. The DTG peaks in the tem­
the presence of NaOH (Fig. 2a), it can be observed that the kaolinite perature range of 400–700 ◦ C were assigned to the dehydroxylation
peak disappeared at around 550 ◦ C. Illite and calcite peaks were peaks for kaolinite and illite, and the decarbonation peak for calcite
diminished at around 630 and 720 ◦ C, respectively. The relatively low

Fig. 1. BSE images and EDS of precursor prepared via alkali fusion, (a–c) control, (d–f) 5 % NaOH, (g–j) 25 % NaOH sample, (k) particle size distribution of the
samples (differential distribution).

4
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

Fig. 2. High-temperature X-Ray Diffraction patterns of (a) control and (b) 10 % NaOH dosage samples. K: Kaolinite, Q: Quartz, C: Calcite, I: Illite, M: Mullite.
N: Nepheline.

Fig. 3. Derivative thermogravimetric (DTG) plots for clay mixed with different NaOH dosages.

based on the literature data [17,59]. Worth to note that illite typically 3.3. Effect of alkali fusion on aluminosilicate
exhibits a broad dehydroxylation temperature range from 450 to 700 ◦ C
[15], and thus, it is difficult to separate it from the decarbonation of 3.3.1. Effect on the crystal structure of raw materials for different NaOH
carbonates, given that the decarbonation temperature of calcium car­ dosage
bonate usually ranges from 200 ◦ C to 800 ◦ C [60], where calcite de­ The XRD spectra of raw, control and alkali fused calcined clays are
composes within the range from 600 ◦ C to 800 ◦ C [47]. Regardless, presented in Fig. 4. The calcination was performed at 600 ◦ C in a muffle
considering the in-situ XRD observation (Fig. 2), the illite and calcite furnace. As observed from the XRD, quartz, illite, calcite, and kaolinite
peaks were marked separately in Fig. 3. Interestingly, after adding 5 % were the dominant phases present in the raw clay mix. The clay was used
NaOH, the dehydroxylation peaks for kaolinite and illite were merged. without drying or removing organic matter for this study. During
Additionally, the dehydroxylation started at a lower temperature (below calcination, organic matters were decomposed, and the sample was fully
300 ◦ C) and was nearly completed by 600 ◦ C. With the addition of 10 or dehydrated. Consequently, the control samples showed a slightly
15 % NaOH, the combined dehydroxylation peak further shifted to a increased intensity for quartz and illite after calcination at 600 ◦ C. The
lower temperature of around 400 ◦ C. With 20 or 25 % of NaOH addition, peak intensity for kaolinite and illite was slightly reduced in the control
there was an additional peak formed at around 750 ◦ C. This new peak batch.
was attributed to the decarbonation of sodium carbonate, which might For the alkali fused batches, the peak intensities of illite and quartz
have formed due to the presence of excess NaOH. Important to note that were reduced with increasing NaOH dosage, which was also evident
the initial calcite decarbonation peak in the range from 600 to 700 ◦ C from the Rietveld quantification (supplementary Figure S1). Such
remained mostly unaltered due to the addition of NaOH. reduced peak intensities were attributed to the formation of the amor­
phous aluminosilicate phase and sodium silicate. However, for a NaOH
dosage of more than 15 %, formation of nepheline was observed.
Nepheline is a crystalline sodium aluminosilicate phase, and previous
studies reported that this phase has a negligible reactivity [43,61].
Therefore, the formation of this phase can negatively affect the

5
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

silicates or aluminosilicates exhibit a broad absorption between 800


cm− 1 and 1200 cm− 1 which corresponds to the asymmetrical stretching
vibration (ν3) of the Si–O bond present in silicate [62]. The peaks rep­
resenting silicate units with various degrees of polymerization (Qn, n =
0, 1, 2, 3, 4) are overlapped in the 800 to 1200 cm− 1 range and
deconvolution is required to separate those peaks. In general, peak at a
lower wavenumber indicates lesser degree of polymerization of the
aluminosilicate or silicate network. From Fig. 5, it can be observed that
the Si–O bond present in the control batch shows a broad peak at around
1000 cm− 1. For the alkali fused batches, with increasing dosage of
NaOH, the ν3 band shifted to a lower wavenumber. This indicates that
the alkali fusion process reduced the overall silicate polymerization of
the calcined clays, and, therefore, confirms that the added Na+ acted as a
network modifier in this case. Accordingly, similar to the previous
observation [43], this finding confirms that co-calcining clays with
NaOH increased the number of non-bridging oxygen (NBO) sites in the
final aluminosilicate or silicate network. Such, an increase in the NBO is
expected to increase the reactivity of the amorphous aluminosilicate
phase [43,63,64]. Worth to note, the FTIR peak at 1440 cm− 1 indicates
C–O bending vibration and can be observed in all the specimens. This
shows that calcination at 600 ◦ C for 60 min was insufficient for the
complete decomposition of CaCO3.
Fig. 4. X-ray powder diffraction of the sample with and without calcination
and different alkali hydroxide dosages. Here, I: Illite (PDF #00-002-0056), K:
3.3.3. Observations from the 29Si NMR spectra
Kaolinite (PDF #01-078-2110), N: Nepheline (PDF #00-002-0637), Q: Quartz 29
Si NMR spectra of calcined clays are presented in Fig. 6. The peaks
(PDF #00-002-0471), Cc: Calcite (PDF#00-005-0586), S: Sodium silicate (PDF
#04-015-9090). were assigned to different silicate species based on the literature
[65–68]. For the control sample, the major peaks at – 94 ppm, − 100
ppm, and − 108.4 ppm corresponding to Q3(1 A l) i.e., [Si(OSi)3(Al)], Q3
reactivity of calcined clays. The above finding suggests that with the
i.e., [Si(OSi)3(OH)], and Q4 i.e., [Si(OSi)4], respectively, are originated
addition of an appropriate amount of NaOH, it is possible to convert the
from illite [68]. The Q4 at − 108.4 ppm also indicates the presence of
typical inter phase present in common clays (i.e., quartz in this case) to
quartz. The minor peaks at − 91.4, − 95.6, and − 100 ppm were assigned
reactive silicate or aluminosilicate phase.
to the amorphous aluminosilicate formed by the calcination of kaolinite,
which is also known as metakaolin [67]. The alkali fused clay with 10 %
3.3.2. Observations from fourier transformed infrared (FTIR) spectra
NaOH showed significantly different NMR spectra compared to the
Fig. 5 illustrates the FTIR spectra of raw clay, calcined clay, and al­
control batch. Specifically, the 10 % NaOH batch showed reduced
kali fused clay with different doses of NaOH. The FTIR spectra for

Fig. 5. FTIR spectra of the samples with and without calcination and different Fig. 6. 29Si NMR spectra of a) control, b) 10 % NaOH, and c) 20 %
alkali hydroxide dosages. NaOH specimen.

6
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

intensity of the peaks at around − 94 ppm, − 100 ppm, and − 108.4 ppm, deviated from the amounts of NaOH (%) that were added. Such devia­
all of which correspond to uncalcined illite [68], while the last peak also tion is expected because the amounts of NaOH (mass %) added were
represents quartz. The 10 % NaOH alkali fused batch showed new peak calculated based on the weight of raw clay. The clay samples are ex­
formation at around − 65 to − 70 ppm and – 85 ppm. Both of these peak pected to lose weight due to the clay dehydroxylation and decomposi­
locations were assigned to the formation of sodium silicate (water glass) tion of organic content, resulting in a higher dosage of Na2O in the final
[66]. Additionally, the peak intensity at around − 90 ppm was increased. SCM compared to the initially added amounts. It is interesting to note
This peak was assigned to Q4 (3 A l) i.e., [Si(OSi) (3 A l)], which can that less than one-fourth of the added dosage of Na2O was released from
originate from amorphous aluminosilicate [65,67] and/or nepheline the alkali-fused clays after the first day of dissolution (Fig. 8b) and the
[69]. For the alkali fused clay with 20 % NaOH, the Q4 peak was shifted leaching rate is approximately proportional to the NaOH dosages. After
from − 108 ppm to − 104 ppm. Intensities of the peaks corresponding to the first day of the dissolution test, the leaching rate of the remaining
Q1 (~80 ppm), Q2 (~85 ppm), and Q4 (3 A l) (~96 ppm) were increased, Na2O was significantly reduced for all of the alkali-fused clays. This
which was attributed to the increased formation of sodium silicate and indicates that the majoring of the Na2O added for the alkali fusion was
amorphous aluminosilicate phases. Overall, the alkali fused clays chemically bound within the structure of alkali-fused clays and was not
showed the presence of 29Si NMR peaks at the lower range indicating a readily extracted.
reduced polymerization of these samples compared to the control batch. The mineralogical alterations of the alkali-fused clays before and
after dissolution in pH 13.5 solution were evaluated using XRD patterns.
The XRD patterns (Fig. 9a) of the control batch (calcined without NaOH)
3.4. Performance of alkali-fused clays as SCMs before and after the dissolution test remained nearly the same. It
matches the relatively low dissolution of Al and Si from this sample as
3.4.1. Dissolution of Al and Si observed in ICP measurements and confirms that the Al and Si dissolu­
The dissolution of Al and Si is a key factor in determining the tion primarily occurred from the amorphous phase present in the
pozzolanic properties of SCMs [70,71]. Studies showed that in the re­ calcined clays. In the case of 5 % NaOH containing alkali-fused clays
action of a silicate-rich SCM in hydrating OPC blend, the dissolution of (Fig. 9b), the before and after XRD patterns were nearly identical to the
silicate species from the SCM particles is the initial rate-controlling step control batch. Therefore, the increase in Al and Si dissolution as
[72,73]. Fig. 7 represents the dissolution of Al and Si from the clay observed in Fig. 7 can be attributed to increased amorphous content in
samples with and without NaOH dosages after 1, 7, and 14 days. The Al these clays. In the case of 15 % and 25 % NaOH containing alkali-fused
dissolution (Fig. 7a) increased by approximately 68 % and 75 % after 14 clays, two new mineral phases were formed after calcination, these are
days for 15 % and 20 % NaOH batches, respectively, compared to the nepheline and sodium silicate. As observed from the XRD patterns
control specimen. The maximum Al dissolution was obtained from 20 % (Fig. 9 c and d), both sodium silicate and nepheline intensity got reduced
NaOH-containing batch (148.4 mg/l). The Si dissolution (Fig. 7b) also with increasing dissolution duration. While previous studies reported
improved with the increase in NaOH content with time. The 20 % dosage that nepheline is non-reactive, the XRD patterns obtained in this work
of the NaOH-containing batch showed the maximum Si dissolution after confirm that nepheline can get dissolved in pH 13.5 releasing additional
14 days (336 mg/l) which is approximately 68 % higher than the control Al, Si, and Na into the matrix. Therefore, for the alkali-fused clays with a
specimen. The slight reduction of Al and Si dissolution for the 25 % higher dosage of NaOH (15 % and 25 %), the dissolution of Al and Si
NaOH containing sample was attributed to the formation of crystalline originates from both the amorphous and crystalline (nepheline, sodium
phases with relatively low reactivity (e.g., nepheline). Regardless, silicate) phases. Comparing the XRD patterns (Fig. 9) with the Na2O
considering the standard deviation (shown as error bars), the dissolu­ measurements from XRF (Fig. 8), it is apparent that around one-fourth of
tions of Al and Si increased with an increase in NaOH dosages. The the added NaOH was bound in the crystalline phases (sodium silicate
amorphization of the clay components, specifically illite in the presence and nepheline), which got nearly fully dissolved within 7 days in pH
of high alkali content (Fig. 4), depolymerized aluminosilicate network 13.5 solution. The remaining NaOH was bound in the amorphous
(Fig. 5), and the sodium silicate formation was attributed to the aluminosilicate, which showed slower leaching of NaOH and was not
enhanced Al and Si dissolution of the samples. detectable in XRD patterns.

3.4.2. Composition and characteristics of alkali-fused clays after 3.4.3. Rapid, relevant, and reliable (R3) method and portlandite
dissolution consumption
The Na2O (mass %) present in the calcined clays before and after The rapid, relevant, and reliable (R3) test is a well-known method to
dissolution in pH 13.5 solution was measured using XRF to understand if represent the pozzolanic activity of a wide range of SCMs [74–76]. Using
the added NaOH is only present as an additive or bound in the SCM. The this test method and its modified versions, the SCMs can be categorized
XRF measurements are shown in Fig. 8. As observed from Fig. 8a, the into inert, pozzolanic, highly pozzolanic, and more hydraulic based on
actual amounts of Na2O (mass %) present in the alkali fused samples

Fig. 7. Dissolution of (a) Al and (b) Si of the alkali fused clays with time.

7
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

Fig. 8. (a) Na2O (mass %) present in alkali fused clays after different exposure durations in pH 13.5 solution and (b) Na2O leaching over time.

Fig. 9. X-ray diffraction patterns of the solid samples of (a) control, (b) 5 % NaOH, (c) 15 % NaOH, (d) 25 % NaOH dosage, before and after dissolution experiments.
Here I: Illite (PDF #00-002-0056), N: Nepheline (PDF #00-002-0637), N’: sodium aluminum silicate hydrate (N-A-S-H) gel, Q: Quartz (PDF #00-002-0471), Cc: 770
Calcite (PDF#00-005-0586), S: Sodium silicate (PDF #04-015-9090).

the total heat release and portlandite consumption data [77,78]. integrating the derivative of thermogravimetric (DTG) peak in the
Fig. 10a represents the total heat release over 168 h of calcined clays as temperature range of 400–500 ◦ C [47], which was then deducted from
obtained using the R3 method. The 15 % NaOH (191.58 J/g) and 25 % the initial amount to determine the Ca(OH)2 consumption. Fig. 10b
NaOH (192.37 J/g) batches had similar heat releases, which were higher represents the quantity of portlandite that was consumed after 7 days in
than all other samples. With the increase in the NaOH content, the heat g per 100 g of solid. It is evident that the relation between portlandite
release also increased. All these clays fall into the ‘Pozzolanic, less consumption and NaOH dosages is linear. The 15 %, 20 %, and 25 %
reactive’ category as their total heat is more than 120 J/g but less than batches consumed around 55–65 % of the total portlandite within 7
370 J/g, according to the literature [78]. days. Considering the error bars, the dissolution of Al and Si with
The amount of Ca(OH)2 present in the samples was determined by different dosage of NaOH fit linearly (Fig. 7a), which matches the

8
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

Fig. 10. (a) Total heat release from R3 test and (b) portlandite consumption after 168 h.

portlandite consumption trend as shown in Fig. 10b. Therefore, the hydration [79]. It is important to note that, for the higher dosages of
higher dissolution of Al and Si enhanced the consumption of more NaOH specimens, sodium silicate formed which is evident from Fig. 4.
portlandite and formed calcium silicate hydrate (C–S–H) [48,50] and in Sodium silicate is known to contribute to the strength gain of a
the presence of Al, calcium aluminum silicate hydrate (C-A-S-H), indi­ cementitious matrix [80]. However, the amount of sodium silicate was
cating higher pozzolanic reactivity. not significantly large enough to render a positive impact on strength
development at the early stage.
3.4.4. Compressive strength and modified strength activity index (SAI) The modified SAI of all the alkali fused batches was in the range of
Fig. 11 represents the compressive strength and modified SAI of 75 %–109 %, implying that these samples satisfy the minimum 75 %
different batches after 7, 14, 28, and 56 days of curing. The compressive limit specified by ASTM C618 [76]. The modified SAI of the alkali-fused
strength and modified SAI of the control batch was around 15 MPa and batch containing 25 % NaOH was lower compared to that of the 20 %
50 %, respectively, after 7 days of curing. After 28 days of curing, the NaOH batch after 7 days of curing. Although the dissolution of Al and Si
control batch had a compressive strength and modified SAI of around 26 were almost in the similar range of these two samples, due to the for­
MPa and 70 %, respectively. Therefore, this low-grade calcined clay mation of crystalline nepheline observed in section 3.4.1, the initial
does not satisfy the SAI requirements of ASTM C618. On the other hand, strength, as well as the modified SAI, was less than that of the 20 %
all of the alkali-fused batches showed relatively higher strength. The NaOH sample.
maximum compressive strength was attained 34 MPa (15 % NaOH) after
28 days of curing. At this curing age, all the alkali fused clay batches 3.4.5. Heat of hydration
showed strength in a similar range (30–34 MPa), indicating their similar The effect of the alkali-fused SCM on the hydration of the OPC system
performance. For the higher NaOH dosages specimens, the strength was was investigated and presented in Fig. 12. It is evident from Fig. 12a that
similar or less than the 5 % NaOH. This could be attributed to the in spite of having a higher amount of free Na2O in the system, the initial
presence of free Na2O in the cement matrix (Fig. 8). The presence of a dissolution of the cement grains was not higher than that of the pure
high quantity of Na2O in the system can have a negative impact on OPC samples. The second peak from 3 to 30 h, representing precipitation
strength development due to the decrease in internal relative humidity, of the reaction product, was also very similar for all of the samples.
which hinders further hydration reaction [79]. It was also found that the However, 5 % NaOH and 10 % NaOH showed a broader peak in this
solution only remains in small pores, which also limits the degree of range, confirming their enhanced reactivity. Fig. 12b exhibits the total

Fig. 11. (a) Compressive strength of the SCM prepared with 20 % replacement of the OPC by calcined low-grade clay samples with and without alkali hydroxide
dosages and (b) modified strength activity index (SAI) over different curing duration.

9
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

Fig. 12. (a) Heat flow and (b) total heat release after 168 h from the OPC paste sample with or without alkali-fused SCMs.

heat release per gram of OPC sample. The maximum total heat was
released from 5 % NaOH (358 J/g). However, OPC, 5 %, 10 %, and 15 %
NaOH specimens had similar total heat release after 168 h. On the other
hand, the control sample showed significantly lower heat release, indi­
cating relatively lower reactivity. These results are in agreement with
the compressive strength results after 7 days.

3.4.6. Mineralogy of the hydration reaction products


The mineralogy of the paste samples prepared with alkali fused clays
as SCM (20 % replacement of OPC) with water to binder ratio of 0.485
(same as the mortar sample) is illustrated in Fig. 13 after 28 days of
curing. The pure OPC sample had a higher quantity of portlandite
compared to the batches containing alkali-fused clay. Due to the
pozzolanic properties, calcined clays consumed portlandite to form
additional calcium aluminum silicate hydrate (C-A-S-H). Consequently,
the portlandite peak intensity was reduced for the samples containing
alkali-fused clays. The 5 % NaOH-containing sample showed the lowest
portlandite peak, which indicates its superior pozzolanic property in the
OPC system. The decrease in the quantity of the portlandite is evident
from supplementary Figure S2. As a result, the application of 5 % NaOH
alkali fused clay as SCM showed the lowest amount of portlandite, even
though the consumption of portlandite followed a linear trend with the
increased NaOH dosage for R3 samples (Fig. 10b). In addition to por­
tlandite, calcium silicate hydrate (C–S–H), calcite (CaCO3), ettringite
[Ca6Al2(SO4)3(OH)12⋅26H2O], and a trace amount of monosulfate
[Ca4Al2O6(SO4)⋅14H2O] [81] and unreacted alite (C3S) and belite (C2S)
were also present in the system [82]. The presence of reactive Al from
the calcined clay favored the formation of AFm-type phases [21,82,83].
The quartz (SiO2) was present in the raw material, and due to its inert
nature, it remained unchanged in the paste samples after 28 days of
curing [14].
To understand if alkali-fused clays would alter the composition of
C–A-S–H when used as an SCM, the SEM-EDS data were also obtained
from 28 days of cured OPC paste samples with 20 % alkali-fused clays as
Fig. 13. X-ray diffraction patterns of the cement paste samples after 28 days of
SCM (Fig. 14). Based on the EDS patterns, it was observed that the alkali
curing containing SCM with and without alkali fusion of different dosages of
uptake by the C–A-S–H was low. The Na/Ca atomic ratio was further NaOH. Here, E: ettringite (PDF #00-041-1451), M: monosulfate, P: portlandite
determined based on the EDS data obtained from 25 different data (PDF #00-044-1481), Q: quartz (PDF #00-002-0471), C: calcium silicate hy­
points for each sample. It was found that the Na/Ca ratio varies from drate (C-S-H)/ calcium aluminum silicate hydrate (C-A-S-H), Cc: calcite
0.002 to 0.04 for the 5 % NaOH-containing sample and from 0.02 to 0.08 (PDF#00-005-0586), A: alite (PDF 96-901-6126), B: belite (PDF #96-
for the 25 % NaOH-containing sample. This confirms that the utilization 154-6029).
of alkali-fused clays as SCM does not significantly alter the C–A-S–H
composition in the OPC-based system. 3.5. Role of alkali-fused clays in alkali silicate reactivity

The expansion due to the alkali-silica reaction is expressed in terms


of change in length shown in Fig. 15. After 16 days of exposure

10
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

Fig. 14. BSE images with EDS of (a–c) control; (d–g) 5 % NaOH and (h–j) 25 % NaOH alkali fused SCM containing sample after 28 days of hydration.

conditions, it was found that the control batch had the lowest rate of the role of available alkali content in SCM has a pessimum effect on the
expansion. When compared with the OPC sample (which does not ASR, that is, high alkali-SCM can also show superior resistance to ASR
contain calcined clays), the control batch showed a percentage reduc­ when used in higher replacement levels [84,85]. Therefore, additional
tion in expansion by 99 %. For the alkali fused batches, the expansion studies are required to assess the ASR performance of mortar bars with
was higher compared to the control batch (which contains low-grade different percent replacements of alkali-fused calcined clays.
calcined clay). The expansion after 16 days of exposure to 1 N NaOH
solution at 80 ◦ C decreased 98 %, 50 % and 7 % for Control, 5 % NaOH 4. Discussions
and 10 % NaOH samples, respectively compared to OPC specimen.
However, the samples expanded 24 % more than that of the OPC sample The co-calcination of aluminosilicates in presence of alkali hydroxide
for 20 % NaOH batch (Figure S3). OPC and alkali-fused batches with helps depolymerize the aluminosilicate network and increases non-
more than 5 % NaOH showed an expansion rate higher than 0.10 %, bridging oxygen (NBO) per tetrahedra [43]. Through a variety of
indicating potentially deleterious expansion in field performance. The experimental techniques, the presented study confirmed that
expansion of the control and 5 % NaOH specimens after 16 days of co-calcining clay with NaOH alters the resultant aluminosilicate phases
exposure was within the limit (less than or equal to 0.1 %). With increase compared to the control batch. Specifically, such alkali fusion reduced
in NaOH content in the alkali fused SCM, free Na2O leaching increased the polymerization of aluminosilicates (as observed from FTIR and 29Si
(Fig. 8b) and consequently the expansion due to alkali silica reaction NMR) and increased the Al and Si dissolutions. Both factors can
augmented in the case of the 10 % and 20 % sample. Important to note, contribute to the enhanced reactivity of the alkali-fused clay. The

11
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

respectively, for 20 % NaOH sample after 14 days compared to


the control sample.
III. The application of alkali fused clays with 5 % and 10% NaOH
dosages as SCM improved the ASR resistance when compared to
the conventional OPC sample.

Overall, this experimental work shows that the reactivity of low-


grade clay can be enhanced by co-calcining (i.e., alkali thermal
fusion) it with alkali hydroxides. However, the suitable alkali dosage is
likely to be dependent on the common clay minerology.

CRediT authorship contribution statement

Ishrat Baki Borno: Data curation, Formal analysis, Investigation,


Writing – original draft, Writing – review & editing. Nithya Nair:
Conceptualization, Formal analysis, Funding acquisition, Investigation,
Project administration, Resources, Supervision, Visualization, Writing –
review & editing. Warda Ashraf: Formal analysis, Writing – review &
editing.

Fig. 15. Comparison of % expansion due to alkali-silica reaction (ASR) of alkali


fused clay used as SCM. Declaration of competing interest

alkali-fused clays with different dosages of NaOH, when used as SCM, The authors declare that they have no known competing financial
attained at least 75 % of the compressive strength of OPC at both 7 and interests or personal relationships that could have appeared to influence
28 days and therefore satisfy the criteria of the modified Strength Ac­ the work reported in this paper.
tivity Index (SAI) (Fig. 11b). The alkali fused clays containing 5 % and
10 % NaOH, when used as SCM, also improved the ASR resistance Data availability
compared to the control batch. Important to note, for most of the per­
formance parameters, performance improved with an increased dosage Data will be made available on request.
of NaOH up to 15 % by weight. After this level, the performance
degraded due to the re-crystallization of the phases. Acknowledgement
One of the interesting observations of this study is the role of NaOH
on the dehydroxylation temperature of clay minerals (Fig. 3). The in-situ This work was conducted with funding support from the US Defense
XRD (Fig. 2) and DTG (Fig. 3) plots revealed that the presence of NaOH Advanced Research Projects Agency (DARPA, award
lowered the temperature required for dehydroxylation and merged the #W911NF2010308) at the University of Texas at Arlington. All opin­
dehydroxylation temperature peaks for kaolinite and illite. If adopted on ions, findings, and conclusions or recommendations expressed in this
an industrial scale, such a technique has the potential to (i) enable material are those of the authors and do not necessarily reflect the views
increased use of common clays with mixed mineral contents by enabling of the sponsor. Dr. Andre Sutrisno, NMR Spectroscopist, School of
calcination at specific temperatures with superior reactivity and (ii) Chemical Sciences, University of Illinois at Urbana-Champaign is
reduce the energy requirement (thus, carbon footprint) of calcined clays gratefully acknowledged for his help and guidance regarding the NMR
by reducing the calcination temperature. Most importantly, the pathway measurements presented in this article.
of utilizing low-grade clays can ensure an abundant supply of SCM to
enhance the sustainability and durability of Portland cement-based
Appendix A. Supplementary data
composites throughout the world. However, the use of NaOH is also
likely to increase the cost and energy requirements of the process.
Supplementary data to this article can be found online at https://doi.
Therefore, detailed technoeconomic and environmental impact assess­
org/10.1016/j.cemconcomp.2023.105417.
ments are required before suggesting if such an approach of using low-
grade clays via alkali-fusion is industrially feasible.
References
5. Conclusions
[1] M.-S. Low, Material Flow Analysis of Concrete in the United States, Master of
Science in Building Technology, Massachusetts Institute of Technology, 2005.
In this study, locally available low-grade clay was utilized to produce [2] J.S. Damtoft, J. Lukasik, D. Herfort, D. Sorrentino, E.M. Gartner, Sustainable
SCMs via alkali fusion. The clay samples were co-calcined with different development and climate change initiatives, Cement Concr. Res. 38 (2008)
115–127.
dosages of NaOH at 600 ◦ C. The followings are the concluding remarks [3] K. Jafari, J. Yoon, R. Tokpatayeva, J. Olek, F. Rajabipour, Surfactant-assisted
of the presented study. purification of an impure kaolinite clay to improve its pozzolanic reactivity in
concrete, J. Mater. Civ. Eng. 34 (2022) 04022094.
[4] M.C.G. Juenger, R. Snellings, S.A. Bernal, Supplementary cementitious materials:
I. Dehydroxylation at 600 ◦ C in the presence of NaOH helped new sources, characterization, and performance insights, Cement Concr. Res. 122
transform a portion of inert quartz to reactive silica, which is in (2019) 257–273.
good agreement with the findings from 29Si spectra, FTIR, and [5] K.-H. Yang, Y.-B. Jung, M.-S. Cho, S.-H. Tae, Effect of supplementary cementitious
materials on reduction of CO2 emissions from concrete, J. Clean. Prod. 103 (2015)
XRD. 774–783.
II. The presence of alkali helped depolymerize the aluminosilicate [6] K. Ezziane, E.-H. Kadri, A. Bougara, R. Bennacer, Analysis of mortar long-term
network and reduced the dehydroxylation temperature. This strength with supplementary cementitious materials cured at different
temperatures, ACI Mater. J. 107 (2010).
overall process contributed to the higher dissolution of reactive
[7] E. Aprianti, P. Shafigh, S. Bahri, J.N. Farahani, Supplementary cementitious
Al and Si. The Si and Al dissolution improved by 68 % and 76 %, materials origin from agricultural wastes–A review, Construct. Build. Mater. 74
(2015) 176–187.

12
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

[8] I. Diaz-Loya, M. Juenger, S. Seraj, R. Minkara, Extending supplementary [38] O.O. Ltaief, S. Siffert, C. Poupin, S. Fourmentin, M. Benzina, Optimal synthesis of
cementitious material resources: reclaimed and remediated fly ash and natural faujasite-type zeolites with a hierarchical porosity from natural clay, Eur. J. Inorg.
pozzolans, Cem. Concr. Compos. 101 (2019) 44–51. Chem. (2015) 4658–4665, https://doi.org/10.1002/ejic.201500537.
[9] F. Massazza, Pozzolana and pozzolanic cements, Lea’s Chem. Cement and Concrete [39] C. Belviso, F. Cavalcante, G. Niceforo, A. Lettino, Sodalite, faujasite and A-type
4 (1998) 471–631. zeolite from 2:1dioctahedral and 2:1:1 trioctahedral clay minerals. A singular
[10] N. Saladi, W. Ashraf, Ground and Sieved Bio Ash versus Coal Fly Ash : Comparative review of synthesis methods through laboratory trials at a low incubation
Analysis of Pozzolanic Reactivity, vol. 32, 2021, pp. 1–13, https://doi.org/ temperature, Powder Technol. 320 (2017) 483–497, https://doi.org/10.1016/j.
10.1061/(ASCE)MT.1943-5533.0003443. powtec.2017.07.039.
[11] I.B. Borno, W. Ashraf, R.I. Khan, A. Tahsin, Multi-technique approaches to evaluate [40] T. Wajima, S. Onishi, Alkali fusion of waste perlite dust to synthesize faujasite
the pozzolanic properties of alkali-rich calcined kaolinite and bentonite clay zeolite using a rotary kiln, Int. J. Chem. Eng. Applic. 10 (2019) 184–188, https://
blends, Adv Civ Eng Mater 11 (2022) 20210156, https://doi.org/10.1520/ doi.org/10.18178/ijcea.2019.10.6.766.
ACEM20210156. [41] T. Wajima, K. Yoshizuka, T. Hirai, Y. Ikegami, Synthesis of zeolite X from waste
[12] K. Scrivener, F. Martirena, S. Bishnoi, S. Maity, Calcined clay limestone cements sandstone cake using alkali fusion method, Mater. Trans. 49 (2008) 612–618,
(LC3), Cement Concr. Res. 114 (2018) 49–56, https://doi.org/10.1016/j. https://doi.org/10.2320/matertrans.MRA2007250.
cemconres.2017.08.017. [42] F. Bullerjahn, M. Zajac, J. Pekarkova, D. Nied, R. Global, D.H. Ag, Novel SCM
[13] S.E. Schulze, J. Rickert, Suitability of natural calcined clays as supplementary produced by the co-calcination of aluminosilicates with dolomite, Cement Concr.
cementitious material, Cem. Concr. Compos. 95 (2019) 92–97. Res. 134 (2020) 106083, https://doi.org/10.1016/j.cemconres.2020.106083.
[14] R. Kaminskas, R. Kubiliute, B. Prialgauskaite, Smectite clay waste as an additive for [43] I.B. Borno, W. Ashraf, Effects of co-calcining kaolinite-rich clay blends with alkali
Portland cement, Cem. Concr. Compos. 113 (2020) 103710. and alkali earth metal hydroxides, Appl. Clay Sci. 231 (2023) 106742.
[15] T. Hanein, K.-C. Thienel, F. Zunino, A.T.M. Marsh, M. Maier, B. Wang, M. Canut, [44] R.K. Mishra, E. Zurich, Comprehensive understanding of grinding aids, ZKG Int. 6
M.C.G. Juenger, M. Ben Haha, F. Avet, Clay calcination technology: state-of-the-art (2014) 28–39.
review by the RILEM TC 282-CCL, Mater. Struct. 55 (2022) 3. [45] J.-B.M. Dassekpo, L. Miao, J. Bai, Q. Gong, N.N. Shao, Z. Dong, F. Xing, J. Ye, Phase
[16] R. Fernandez, F. Martirena, K.L. Scrivener, The origin of the pozzolanic activity of dissolution and improving properties of completely decomposed granite through
calcined clay minerals: a comparison between kaolinite, illite and montmorillonite, alkali fusion method, Cem. Concr. Compos. 127 (2022) 104407.
Cement Concr. Res. 41 (2011) 113–122. [46] Standard Test Methods for Measuring the Reactivity of Supplementary
[17] A.Z. Khalifa, Ö. Cizer, Y. Pontikes, A. Heath, P. Patureau, S.A. Bernal, A.T. Cementitious Materials by Isothermal Calorimetry and Bound Water Measurements
M. Marsh, Advances in alkali-activation of clay minerals, Cement Concr. Res. 132 vol. 1, (n.d.). https://doi.org/10.1520/C1897-20.
(2020) 106050. [47] K. Scrivener, R. Snellings, B. Lothenbach, A Practical Guide to Microstructural
[18] R. Jaskulski, D. Jóźwiak-Niedźwiedzka, Y. Yakymechko, Calcined clay as Analysis of Cementitious Materials, Crc Press, Boca Raton, FL, USA, 2016.
supplementary cementitious material, Materials 13 (2020) 4734. [48] M. Mejdi, W. Wilson, M. Saillio, T. Chaussadent, L. Divet, A. Tagnit-Hamou,
[19] H. Du, A. Dixit, S.D. Pang, Potential of marine clay for cement replacement and Investigating the pozzolanic reaction of post-consumption glass powder and the
pozzolanic additive in concrete, in: Calcined Clays for Sustainable Concrete, role of portlandite in the formation of sodium-rich CSH, Cement Concr. Res. 123
Springer, 2020, pp. 57–65. (2019) 105790.
[20] M. Bediako, S.K.Y. Gawu, A.A. Adjaottor, J.S. Ankrah, Early and late strength [49] S. Wild, J.M. Khatib, Portlandite consumption in metakaolin cement pastes and
characterization of portland cement containing calcined low-grade kaolin clay, mortars, Cement Concr. Res. 27 (1997) 137–146.
J. Eng. (2016) 2016. [50] S. Hollanders, R. Adriaens, J. Skibsted, Ö. Cizer, J. Elsen, Pozzolanic reactivity of
[21] A. Alujas, R. Fernández, R. Quintana, K.L. Scrivener, F. Martirena, Pozzolanic pure calcined clays, Appl. Clay Sci. 132 (2016) 552–560.
reactivity of low grade kaolinitic clays: influence of calcination temperature and [51] Y. Wang, P. Suraneni, Experimental methods to determine the feasibility of steel
impact of calcination products on OPC hydration, Appl. Clay Sci. 108 (2015) slags as supplementary cementitious materials, Construct. Build. Mater. 204 (2019)
94–101. 458–467.
[22] N.R. Rakhimova, R.Z. Rakhimov, A.R. Bikmukhametov, V.P. Morozov, A.A. Eskin, [52] Standard Test Methods for Sampling and Testing Fly Ash or Natural Pozzolans for
T.Z. Lygina, A.M. Gubaidullina, Role of clay minerals content and calcite in alkali Use in Portland-Cement Concrete vol. 1, (n.d.). https://doi.org/10.1520/C0311_
activation of low-grade multimineral clays, J. Mater. Civ. Eng. 32 (2020) C0311M-22.
04020198. [53] Standard Test Method for Compressive Strength of Hydraulic Cement Mortars
[23] M. Flegar, K. Ram, M. Serdar, K. Bosnar, K. Scrivener, Durability challenges of low- (Using 2-in. Or [50-mm] Cube Specimens), (n.d.)..
grade calcined clay opposed to high-volume fly ash in general purpose concrete, [54] S. Al-Shmaisani, R.D. Kalina, R.D. Ferron, M.C.G. Juenger, Critical assessment of
Adv Civ Eng Mater 12 (2023) 20220138, https://doi.org/10.1520/ rapid methods to qualify supplementary cementitious materials for use in concrete,
ACEM20220138. Cement Concr. Res. 153 (2022) 106709.
[24] P. Lehmann, B. Leshchinsky, S. Gupta, B.B. Mirus, S. Bickel, N. Lu, D. Or, Clays are [55] Standard Test Method for Determining the Potential Alkali-Silica Reactivity of
not created equal: how clay mineral type affects soil parameterization, Geophys. Combinations of Cementitious Materials and Aggregate (Accelerated Mortar-Bar
Res. Lett. 48 (2021), https://doi.org/10.1029/2021GL095311. Method) 1, (n.d.). https://doi.org/10.1520/C1567-21.
[25] X. Jian, J. Huang, Z. Cai, Effect of alkaline fusion on muscovite decomposition and [56] N. Blouch, K. Rashid, I. Zafar, M. Ltifi, M. Ju, Prioritization of low-grade kaolinite
the vanadium release mechanism from vanadium shale, R. Soc. Open Sci. 5 (2018). and mixed clays for performance evaluation of Limestone Calcined Clay Cement
[26] A.A. Shoppert, I. V Loginova, D.A. Rogozhnikov, K.A. Karimov, L.I. Chaikin, (LC3): multi-criteria assessment, Appl. Clay Sci. 243 (2023) 107080, https://doi.
Increased as adsorption on maghemite-containing red mud prepared by the alkali org/10.1016/j.clay.2023.107080.
fusion-leaching method, Minerals 9 (2019), https://doi.org/10.3390/min9010060. [57] N. Garg, J. Skibsted, Pozzolanic reactivity of a calcined interstratified illite/
[27] S.N. Ishmah, M.D. Permana, M.L. Firdaus, D.R. Eddy, Extraction of silica from smectite (70/30) clay, Cement Concr. Res. 79 (2016) 101–111, https://doi.org/
bengkulu beach sand using alkali fusion method, PENDIPA J. Sci. Educ. 4 (2020) 10.1016/j.cemconres.2015.08.006.
1–5. [58] L. Fernandes, R. Salomão, Preparation and characterization of mullite-alumina
[28] I. Choi, G. Moon, J. Lee, R.K. Jyothi, Alkali fusion using sodium carbonate for structures formed" in situ" from calcined alumina and different grades of synthetic
extraction of vanadium and tungsten for the preparation of synthetic sodium amorphous silica, Mater. Res. (2018) 21.
titanate from spent SCR catalyst, Sci. Rep. 9 (2019) 1–8, https://doi.org/10.1038/ [59] W. Ashraf, Carbonation of cement-based materials: challenges and opportunities,
s41598-019-48767-0. Construct. Build. Mater. 120 (2016) 558–570, https://doi.org/10.1016/j.
[29] R. Goguel, Alkali release by volcanic aggregates in concrete, Cement Concr. Res. 25 conbuildmat.2016.05.080.
(1995) 841–852, https://doi.org/10.1016/0008-8846(95)00075-N. [60] M.I. Haque, I.B. Borno, R.I. Khan, W. Ashraf, Reducing carbonation degradation
[30] F. Wulandari, E.P. Ramdhani, Y. Lailun, A.A. Dawam, D. Prasetyoko, Synthesis of and enhancing elastic properties of calcium silicate hydrates using biomimetic
amorphous mesoporous aluminosilicates from bintan’ s red mud as alumina source, molecules, Cem. Concr. Compos. 136 (2023) 104888.
Indonesian J. Chem. 18 (2018) 580–586, https://doi.org/10.22146/ijc.25184. [61] M.X. Peng, Z.H. Wang, S.H. Shen, Q.G. Xiao, L.J. Li, Y.C. Tang, L.L. Hu, Alkali
[31] T. Wajima, Alkali conversion of waste clay into zeolitic materials using NaOH and fusion of bentonite to synthesize one-part geopolymeric cements cured at elevated
KOH solution, Int. J. Soc. Mater. Eng. Resour. 23 (2018) 30–35. temperature by comparison with two-part ones, Construct. Build. Mater. 130
[32] H. Ma, Q. Yao, Y. Fu, C. Ma, X. Dong, Synthesis of zeolite of type A from bentonite (2017) 103–112.
by alkali fusion activation using Na2CO3, Ind. Eng. Chem. Res. 49 (2010) 454–458. [62] W. Ashraf, J. Olek, Carbonation behavior of hydraulic and non-hydraulic calcium
[33] W. Xuan, Q. Wang, J. Zhang, D. Xia, Influence of silica and alumina (SiO 2 + Al 2 silicates: potential of utilizing low-lime calcium silicates in cement-based
O3) on crystallization characteristics of synthetic coal slags, Fuel 189 (2017) materials, J. Mater. Sci. 51 (2016) 6173–6191, https://doi.org/10.1007/s10853-
39–45, https://doi.org/10.1016/j.fuel.2016.10.081. 016-9909-4.
[34] T. Wajima, K. Sugawara, Material conversion from various incinerated ashes using [63] D. Feng, J.L. Provis, J.S.J. van Deventer, Thermal activation of albite for the
alkali fusion method, Int. J. Soc. Mater. Eng. Resour. 17 (2010) 47–52. synthesis of one-part mix geopolymers, J. Am. Ceram. Soc. 95 (2012) 565–572.
[35] J.E. Shelby, Introduction to Glass Science and Technology, second ed., The Royal [64] X. Ke, S.A. Bernal, N. Ye, J.L. Provis, J. Yang, One-part geopolymers based on
Society of Chemistry 2005, 2005. thermally treated red mud/NaOH blends, J. Am. Ceram. Soc. 98 (2015) 5–11.
[36] S. Nie, R.M. Thomsen, J. Skibsted, Impact of Mg substitution on the structure and [65] X. Gao, Q.L. Yu, H.J.H. Brouwers, Apply 29Si, 27Al MAS NMR and selective
pozzolanic reactivity of calcium aluminosilicate (CaO-Al2O3-SiO2) glasses, dissolution in identifying the reaction degree of alkali activated slag-fly ash
Cement Concr. Res. 138 (2020) 106231. composites, Ceram. Int. 43 (2017) 12408–12419, https://doi.org/10.1016/j.
[37] L. Tosheva, A. Brockbank, B. Mihailova, J. Sutula, J. Ludwig, H. Potgieter, ceramint.2017.06.108.
J. Verran, Micron-and nanosized FAU-type zeolites from fly ash for antibacterial [66] H. Jansson, D. Bernin, K. Ramser, Silicate species of water glass and insights for
applications, J. Mater. Chem. 22 (2012) 16897–16905. alkali-activated green cement, AIP Adv. 5 (2015), https://doi.org/10.1063/
1.4923371.

13
I.B. Borno et al. Cement and Concrete Composites 147 (2024) 105417

[67] H.P. He, J.G. Guo, J.X. Zhu, C. Hu, 29 Si and 27 Al MAS NMR study of the thermal [77] P. Suraneni, J. Weiss, Examining the pozzolanicity of supplementary cementitious
transformations of kaolinite from North China, Clay Miner. 38 (2003) 551–559, materials using isothermal calorimetry and thermogravimetric analysis, Cem.
https://doi.org/10.1180/0009855033840114. Concr. Compos. 83 (2017) 273–278.
[68] W. Luo, T. Fukumori, B. Guo, K. Osseo-Asare, T. Hirajima, K. Sasaki, Effects of [78] P. Suraneni, A. Hajibabaee, S. Ramanathan, Y. Wang, J. Weiss, New insights from
grinding montmorillonite and illite on their modification by dioctadecyl dimethyl reactivity testing of supplementary cementitious materials, Cem. Concr. Compos.
ammonium chloride and adsorption of perchlorate, Appl. Clay Sci. 146 (2017) 103 (2019) 331–338.
325–333, https://doi.org/10.1016/j.clay.2017.06.025. [79] M. Zhang, F. Zunino, L. Yang, F. Wang, K. Scrivener, Understanding the negative
[69] J.F. Stebbins, J.B. Murdoch, I.S.E. Carmichael, A. Pines, Defects and Short-Range effects of alkalis on long-term strength of Portland cement, Cement Concr. Res. 174
Order in Nepheline Group Minerals: a Silicon-29 Nuclear Magnetic Resonance (2023) 107348, https://doi.org/10.1016/j.cemconres.2023.107348.
Study, 1986. [80] X. Dai, Q. Ren, S. Aydin, M.Y. Yardimci, G. De Schutter, Accelerating the reaction
[70] R. Snellings, Solution-controlled dissolution of supplementary cementitious process of sodium carbonate-activated slag mixtures with the incorporation of a
material glasses at pH 13: the effect of solution composition on glass dissolution small addition of sodium hydroxide/sodium silicate, Cem. Concr. Compos. 141
rates, J. Am. Ceram. Soc. 96 (2013) 2467–2475. (2023) 105118, https://doi.org/10.1016/j.cemconcomp.2023.105118.
[71] S. Li, G. Chen, Y. Zhao, Z. Xu, X. Luo, C. Liu, J. Gao, Investigation on the reactivity [81] H. Lee, D. Jeon, H. Song, S.W. Sim, D. Kim, J. Yu, K.H. Cho, J.E. Oh, Recycling of
of recycled brick powder, Cem. Concr. Compos. 139 (2023) 105042. reverse osmosis (RO) reject water as a mixing water of calcium sulfoaluminate
[72] N. Garg, J. Skibsted, Dissolution kinetics of calcined kaolinite and montmorillonite (CSA) cement for brick production, Appl. Sci. 9 (2019) 5044.
in alkaline conditions: evidence for reactive Al (V) sites, J. Am. Ceram. Soc. 102 [82] S. Srivastava, M. Cerutti, H. Nguyen, V. Carvelli, P. Kinnunen, M. Illikainen,
(2019) 7720–7734. Carbonated steel slags as supplementary cementitious materials: reaction kinetics
[73] S.A. Greenberg, Reaction between silica and calcium hydroxide solutions. I. and phase evolution, Cem. Concr. Compos. (2023) 105213.
Kinetics in the temperature range 30 to 85◦ , J. Phys. Chem. 65 (1961) 12–16. [83] W. Ashraf, I.B. Borno, R.I. Khan, S. Siddique, M.I. Haque, A. Tahsin, Mimicking the
[74] F. Avet, R. Snellings, A.A. Diaz, M. ben Haha, K. Scrivener, Development of a new cementation mechanism of ancient Roman seawater concrete using calcined clays,
rapid, relevant and reliable (R3) test method to evaluate the pozzolanic reactivity Appl. Clay Sci. 230 (2022) 106696.
of calcined kaolinitic clays, Cement Concr. Res. 85 (2016) 1–11. [84] Chau Lee, Available Alkalis in Fly Ash and Their Effects on Alkali-Aggregate
[75] R. Snellings, K.L. Scrivener, Rapid screening tests for supplementary cementitious Reaction, Iowa State University, 1986.
materials: past and future, Mater. Struct. 49 (2016) 3265–3279. [85] R. Detwiler, The Role of Fly Ash Composition in Reducing Alkali-Silica Reaction,
[76] J. Yoon, K. Jafari, R. Tokpatayeva, S. Peethamparan, J. Olek, F. Rajabipour, 1997. Illinois.
Characterization and quantification of the pozzolanic reactivity of natural and non-
conventional pozzolans, Cem. Concr. Compos. 133 (2022) 104708.

14

You might also like