You are on page 1of 10

Journal of Cleaner Production 383 (2023) 135439

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

Development of high-strength geopolymers from red mud and blast


furnace slag
Umme Zakira a, *, Kai Zheng b, d, Ning Xie b, **, Bjorn Birgisson c
a
Zachry Department of Civil and Environmental Engineering, Texas A&M University, College Station, Texas, 77843-3136, USA
b
Shandong Provincial Key Laboratory of Preparation and Measurement of Building Materials, University of Jinan, Jinan, 250022, China
c
School of Environmental, Civil, Agricultural, and Mechanical Engineering, University of Georgia, Athens, Georgia, 30602, USA
d
Institute of Highway Engineering, RWTH Aachen University, Aachen, 52074, Germany

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

Handling Editor: Prof. Jiri Jaromir Klemeš The worldwide growing interest in green construction urges for innovation of building materials by reusing the
existing waste resources, especially industrial by-products like red mud, which has been an environmental
Keywords: concern in the last few decades. This paper presents a study that investigated the potential use of red mud as a
Waste reuse raw material in the development of high-strength geopolymers. A Sodium hydroxide (NaOH) and Sodium silicate
Red mud
(Na2SiO3) solution were used as the activation reagent for red mud, and blast furnace slag-derived geopolymers.
High-strength geopolymer
This study adopted two different mixing procedures while activating the raw ingredients. Additionally, Ten
Blast-furnace slag
Silica fume percent (10%) by weight silica fume was introduced as an additive to assess its impact on the formed geo­
Compressive strength polymers. The mechanical and microstructural properties of the geopolymers were characterized based on two
Microstructural characterization types of preparation processes. The synthesized geopolymers achieved extraordinary early compressive strength
as high as 68.8 MPa. A high proportion of red mud, 50 percent by weight, could be utilized as a raw ingredient
achieving a high strength of 66 MPa. The optimum initial molar ratios were reported based on the geopolymer
products’ compressive strength and microstructural performances. Silica fume was found to be an efficient ad­
ditive that enhanced overall performance, ensuring higher proportional use of red mud as a raw material.

1. Introduction applications, including its use as an environmentally friendly cementi­


tious material.
Geopolymers are sustainable cementitious materials that have the The geopolymerization process involves the dissolution of alumino­
potential to replace ordinary Portland cement (OPC), which is respon­ silicate oxides (Al3+ in IV-fold coordination) in alkali poly-silicate,
sible for as much as 8% of the world’s carbon dioxide (CO2) emissions which yields polymeric Si–O–Al bonds and forms three-dimensional
(Andrew, 2018). Their performance is similar to that of OPC, but they semi-crystalline silico-aluminate structures. The general formula of the
have the advantage of reclaiming industrial wastes as a source of silico aluminate structure is Mn[–(Si–O2)z–Al–O]n.wH2O, where M rep­
alumina silicate precursor resources (Hu et al., 2018). While resents one or more alkali metals and z is the degree of polycondensation
manufacturing hydraulic cement can cause from 600 kg to 1200 kg per or polymerization (1, 2, 3, or higher) (Davidovits et al., 1991; Davidovits
metric ton of carbon dioxide (CO2), depending on the efficiency of the et al., 1994). There are three basic forms of geopolymers: poly(sialate),
production and proportion of clinker (Y. Kim et al., 2018; Hanif et al., poly(sialate-siloxo), and poly(sialate-disiloxo) with a Si/A1 of 1, 2, and
2017), the production of geopolymer cement results in carbon dioxide 3, respectively (Davidovits et al., 1991). Davidovits et al. (1994) sug­
(CO2) emissions ranging between 271 kg and 425 kg per metric ton gested molar ratios of SiO2/M2O and SiO2/Al2O3 in the range of 4.0–6.6
(Mclellan et al., 2011). In addition to being sustainable, geopolymers and 5.5–6.5, respectively, for alkaline poly(sialate-disiloxo) with a Si/Al
have advantages over OPC in terms of strength, high-temperature of 3. A similar range of molar ratios, such as M2O/SiO2, 0.2–0.48,
resistance, shrinkage, and acid resistance (Hu et al., 2019). The multi­ SiO2/Al2O3, 3.3–4.5, and H2O/M2O, 10–25, was found suitable in other
fold benefits of geopolymers unlock a notable range of possible studies (Khale and Chaudhary, 2007; Ye et al., 2016). Thus,

* Corresponding author.
** Corresponding author.
E-mail addresses: uz17@tamu.edu (U. Zakira), mse_xien@ujn.edu.cn (N. Xie).

https://doi.org/10.1016/j.jclepro.2022.135439
Received 22 December 2021; Received in revised form 22 November 2022; Accepted 28 November 2022
Available online 2 December 2022
0959-6526/© 2022 Elsevier Ltd. All rights reserved.
U. Zakira et al. Journal of Cleaner Production 383 (2023) 135439

geopolymerization needs an optimum molar proportion of alumina and there are limited studies on the effect of the mixing procedure and the
silicate, along with an alkaline medium, e.g., Na, K- hydroxide, a molar ratio of RM-based geopolymers. Moreover, compressive strength
Na2SiO3 solution that can serve as an alkali activator. Sodium silicate of 55 MPa or higher is designated as high-strength concrete by ACI
and sodium hydroxide solution are commonly used as the alkaline so­ 363R, 2010. In several research works, high-strength geopolymers have
lution, where OH− helps in the dissolution of Si4+ and Al3+ of the been produced with high calcium fine fly-ash-based mortar (Chindap­
feedstock, and Na+ contributes to the crystallization of the geopolymers rasirt et al., 2011), fly ash, and slag-based concrete (Neupane, 2018),
(Murayama et al., 2002). The sodium silicate to sodium hydroxide ratio yet, there are few studies on the development of high-strength geo­
plays a crucial role, as both low and high amounts of Na+ can impact polymers utilizing RM and industrial wastes.
compressive strength. Less Na+ and OH− impede complete dissolution This study aimed to develop high-strength geopolymers using RM.
and polymerization, and a high Na+ content weakens the geopolymer Waste materials, including RM and GGBS, were used as the source of
structure by leaving the excess Na+ residue in the sample (Rowles and alumina and silica, and sodium silicate and sodium hydroxide (NaOH)
O’Connor, 2003). Morsy et al. (2014) reported that the compressive solutions were used as alkaline reagents. The mechanical and micro­
strength of fly ash geopolymers increases as the sodium silicate to so­ structural performances of the RM- and GGBS-based geopolymers were
dium hydroxide ratio increases from 0.5 to 1.0, then decreases as the investigated by conducting a pH test, particle size analysis, x-ray fluo­
ratio decreases from 1 to 2.5. The aluminate silicate precursors of geo­ rescence (XRF), x-ray diffraction (XRD), scanning electron microscopy
polymers can be sourced from natural resources, e.g., metakaolin (SEM), and compressive strength tests. The effect of molar ratios and two
(Duxson et al., 2007), kaolinite (Xu and Van Deventer, 2003), and in­ different mixing procedures have been investigated based on the overall
dustrial by-products, such as fly ash (Jiang et al., 2020; Jiang et al., performances of the synthesized geopolymers. Silica fume was incor­
2020; Y. Wang et al., 2021), waste glass (Zhang et al., 2020; Xiao et al., porated into the optimized mixture to assess its impact on the me­
2020), slag (Xiao et al., 2021; S. Li et al., 2021) and soil waste (Capasso chanical and microstructural properties.
et al., 2020) and can be used effectively with varying alkaline dosages,
activation processes and curing conditions. 2. Materials and methods
Red mud (RM) or bauxite residue is a highly alkaline industrial waste
with an average pH of 11.3 ± 1.0 (Gräfe et al., 2011) that amounts to 2.1. Materials
about 55–65% of the processed bauxite ores in alumina refineries
(Paramguru et al., 2004). According to the US Geological Survey (Na­ Locally sourced RM (Chipin Group, Shandong, China) and
tional Minerals Information Center, 2020), 370 million metric tons of commercially available ground-granulated-blast-furnace slag (GGBS)
bauxite ores were mined globally in 2019, consequently expanding the (Gongyi Longze Water Purification Material Co., LTD) were used as raw
quantity of RM as a by-product to approximately 175 million tons materials for the geopolymer synthesis. The RM was air-dried and pul­
annually. China generates 70 million tons of alumina every year, which verized in a jaw crusher and oven-dried at 80 ◦ C for 12 h before being
represents more than half of the alumina produced globally and results ball-milled at 260 rpm for 3 min clockwise and another 3 min counter-
in an annual output of more than 70 million tons of RM. As massive clockwise. It was then passed through a No. 60 sieve (250 μm) to ensure
quantities of RM cannot be used effectively, China has a stockpile of its homogeneity and fineness. The alkaline reagent was made with
about 600 million tons that takes up large amounts of land and poses commercially available sodium hydroxide (NaOH) flakes (Sigma-
serious environmental risks (Zinoveev et al., 2021; Agrawal et al., 2022). Aldrich Co.) and water glass (WG) (Jiashan Yurui Refractory Material
Currently, the most common practice of managing RM is depositing it in Co., LTD) with 44% (by weight) sodium silicate SiO2: Na2O: H2O = 3.4:
a tailing pond or a landfill, which raises environmental concerns due to 1: 5.6). Commercially available silica fume (SF) (Gongyi Longze Water
its causticity, space consumption, and possible groundwater pollution. Purification Material Co., LTD) was used in combination with RM and
This urges for development of strategies for utilizing and managing GGBS to add active silica to the system.
these wastes. Utilization of red mud in building materials can be a viable
management strategy that may significantly reduce or prevent the
2.2. Methods
storage of the red mud (Liu et al., 2014; Vigneshwaran et al., 2019,
2020; S. Wang et al., 2021). In such a way, the goal of obtaining sus­
The mix proportions of the RM, GGBS slag, sodium silicate, and
tainable and cleaner production of low-carbon construction materials
NaOH used in the geo-polymerization process for this study are provided
can be achieved with the reuse of red mud on a large scale. As RM
in Table 1. Different weight percentages of RM and GGBS were incor­
contains alumina (Al2O3), hematite (Fe2O3), silicates (SiO2), and alkali
porated, keeping the alkaline reagent constant. RM50-ASF was prepared
oxides (Na, Ca, K), it can potentially be used as a raw ingredient in the
by replacing 10-wt percent of the GGBS with silica fume to assess the
development of geopolymers. It typically contains a low percentage of
effect. Two methods were followed to prepare the samples. Method A
alumina and mostly non-reactive silica (Singh et al., 2018). Therefore, it
involved homogenous mixing of dry ingredients (RM, GGBS, and SF) at
is usually pretreated and mixed with additional sources of alumina and
1500 rpm for 30 s, then adding a NaOH solution and mixing for another
active silica such as fly ash, Ground Granulated Blast-furnace Slag
30 s. After a 15-s delay, 196 gm of the sodium silicate was added and
(GGBS), silica fume, rice husk, metakaolin (Ye et al., 2016; Singh et al.,
mixed for 15 s at 1500 rpm and 30 s at 3000 rpm. The ratio of anhydrate
2018; He et al., 2013; S. Li et al., 2018; Hu et al., 2018; Nie et al., 2016).
sodium silicate to sodium hydroxide and water to binder (w/b) was 1.6
Silica fume increases the reactivity of Al3+ in GGBS more effectively
than fly ash and increases the reactivity of Na+ in the alkali reagent (Lee
Table 1
et al., 2016). Calcined feedstocks, e.g., GGBS, fly ash, and metakaolin,
Mix proportions (wt%) used for red mud (RM)-based geopolymers.
contain high amounts of CaO and exhibit higher reactivity during geo­
polymerization than non-calcined ones and enhance microstructural Method Mix ID RM GGBS SF
and mechanical properties of geopolymers (Khale and Chaudhary, Method A RM10-AS 10 90 0
2007). The dissolution of RM or aluminosilicate sources plays a vital role RM30-AS 30 70 0
in the physical and mechanical performance of the final products (Lig­ RM50-AS 50 50 0
RM50-ASF 50 40 10
uori et al., 2017). Several pretreatment strategies have been suggested in RM70-AS 70 30 0
the previous research to activate relevant elements in RM by thermal RM90-AS 90 10 0
(Hairi et al., 2015; Ye et al., 2014), alkali-thermal (Ye et al., 2016), Method B RM50-BS 50 50 0
mechanical activation (Y. Li et al., 2019; Singh et al., 2018) or combi­ RM70-BS 70 30 0
RM90-BS 90 10 0
nation of them (Kumar et al., 2021; Hertel and Pontikes, 2020). Hence,

2
U. Zakira et al. Journal of Cleaner Production 383 (2023) 135439

and 0.26, respectively, for all the mix proportions in method A. In


preparation for NaOH solution, 55 gm of NaOH flakes were mixed in tap
water and cooled at room temperature to adjust 3.5M sodium silicate to
1.2M solution. In Method B, after the first mixing stage, the activator
solutions (solution of 10M NaOH and sodium silicate) were added all
together and mixed for 90 s at 1500 rpm. The ratio of sodium silicate to
sodium hydroxide solution and water to binder was 2.5 and 0.26,
respectively. The details of the mixing processes are shown in Fig. 1.
Specimens with dimensions of 40 mm × 40 mm x 40 mm were
prepared using oiled standard steel molds. The specimens were placed in
a standard curing chamber at 60 ◦ C ambient temperature. They were
then removed from the molds after 24 h of curing and continued to cure
at the same temperature. A Universal Testing Machine (CMT5504) was
used to conduct compressive strength tests after 3 days and 28 days of
curing, according to ASTM C39 (ASTM C39/C39M, 2003) or GBT
50107–2010 (GB/T 50107-2010, 2010) at a constant strain rate of 0.5%
strain rate per minute. The standard error values for compressive
strength were calculated based on 3 replicates of strength tests. The pH
of the RM solution was measured by inserting a pH meter into a solution
of 2 gm of RM and 20 ml of water. The density of the samples was
measured according to ASTM C188 (ASTM C188-16, 2009) and the
Fig. 2. Particle size distribution of red mud (RM) and ground granulated blast-
particle sizes were analyzed by an LS13320 laser particle size analyzer. furnace slag (GGBS).
Powder XRD patterns were collected from samples on a D8ADVANCE
diffractometer, and a JSM-7610F field-emission scanning electron mi­
croscope (SEM) was used to image the morphology of the samples. Table 2
Chemical constituents of red mud (RM), ground-granulated blast-furnace slag
3. Results and discussions (GGBS) and silica fume (SF), wt%.
Components RM GGBS SF
3.1. Physical and microstructural properties of raw materials Fe2O3 34.81 0.29 0.04
SiO2 12.94 34.11 98.08
3
The density and pH of the RM were found to be 2.94 g/cm and Na2O 10.92 0.75 –
10.32, respectively. The high pH in RM results from the presence of Al2O3 – 17.77 0.34
TiO2 5.41
alkali metals. The density of GGBS is similar to that of RM (2.99 g/cm3); – –
CaO 4.2 33.15 0.12
the d50 of RM is 17.24 μm and of GGBS is 10.20 μm. The particles of the SO3 0.49 – 0.42
GGBS are finer than those of RM which have a larger surface area and P2O5 0.32 0.06 0.55
dissolve and react more rapidly in forming the geopolymer gel. Fig. 2 ZrO2 0.18 – –
shows the particle distribution of the raw materials. V2O5 0.16 – –
MgO 0.16 0.34
The X-ray fluorescence (XRF) results presented in Table 2 reveal that

K2O 0.14 – –
RM is rich in Fe2O3, SiO2, and alkali metals, but has no trace of Al2O3. MnO 0.1 – –
GGBS is rich in SiO2, Al2O3, and CaO, which indicates that it is a suitable Cr2O3 0.08 – –
supplement for the preparation of the geopolymer. Silica fume (SF) Cl 0.04 – –
contains SiO2 and a small number of other minerals. Both the XRD
pattern and XRF analysis results showed a similar chemical composition
3.2. Mechanical properties of geopolymer
of the raw materials (Fig. 3 and Table 2). According to Fig. 3, hematite
(Fe2O3) at 33◦ –37◦ (2θ), quartz (SiO2) at 22◦ (2θ), katoite (Ca2.93Al1.97
Table 3 shows the compressive strength of the geopolymers that were
Si0.64O2.56(OH)9.44) at 17◦ –18◦ (2θ) and calcite (CaCO3) at 29◦ (2θ) are
derived from RM and GGBS and activated with a NaOH and sodium
the major crystalline compounds present in the RM. Whereas, A broad
silicate solution. When the proportion of RM reduced, the compressive
bump distributed between 20◦ (2θ) and 36◦ (2θ) can be observed for
strength of the RM-GGBS-based geopolymers increased, except for the
GGBS indicating the existence of amorphous phases. The micrographs of
RM30-AS, which has a higher standard error than other mixtures. High
the RM and GGBS are presented in Fig. 4.
strength was achieved with the addition of up to 70% RM, by ACI 363R,
2010, as shown in Table 3.

Fig. 1. Details of mixing methods A and B used for red mud (RM)-based geopolymers.

3
­
U. Zakira et al. Journal of Cleaner Production 383 (2023) 135439

Table 3
Compressive strength of red mud (RM)-based geopolymers.
Mix ID Initial molar ratio Compressive
Strength (MPa)

SiO2/ SiO2/ Na2O/ CaO/ 3d 28d


Al2O3 Na2O Al2O3 SiO2

RM10- 4.405 4.020 1.096 0.781 68.79 ± 68.62 ±


AS 5.91 1.99
RM30- 5.086 3.031 1.678 0.703 52.29 ± 60.44 ±
AS 6.57 8.03
RM50- 6.312 2.316 2.726 0.605 61.21 ± 65.70 ±
AS 2.02 2.02
RM50- 9.372 2.778 3.374 0.418 67.85 ± 67.29 ±
ASF 0.90 1.81
RM70- 9.172 1.774 5.171 0.479 45.97 ± 51.82 ±
AS 1.57 3.35
RM90- 23.472 1.349 17.395 0.309 12.05 ± 8.76 ±
AS 0.13 0.27
RM50- 4.974 2.164 2.299 0.768 19.06 ± n/a
BS 0.89
Fig. 3. XRD patterns of red mud (RM) and ground-granulated blast-furnace slag RM70- 6.942 1.557 4.459 0.633 14.33 ± n/a
(GGBS) (S, H, and C denote peaks of Tridymite, Hematite and calcite, BS 1.20
respectively). RM90- 16.782 1.100 15.259 0.433 3.99 ± n/a
BS 0.78

The initial molar ratios of Na2O/Al2O3 and SiO2/Al2O3 increased, N.B: n/a = not available (not tested).
and SiO2/Na2O and CaO/SiO2 decreased with the addition of the higher
amount of RM. High strength, up to 68.6 MPa, was achieved with RM-
and GGBS-based geopolymers. It is clear from Table 3 and Fig. 5 that the
compressive strength declined from geopolymers RM10-AS to RM90-AS,
correlating with the molar ratio of the elemental components of the raw
materials. Fig. 5 shows that the synthesized geopolymers attained 90%
and more of 28-day strength within 3 days of curing.
It is evident in Fig. 5 that a high strength of 65.7 MPa has been
achieved with Na2O/Al2O3 and SiO2/Al2O3 ratios of 2.7 and 6.3 when
50 wt percent of RM is incorporated. The addition of GGBS increases the
SiO2, Al2O3, and CaO content and maintains the balance of the molar
ratios. Fig. 6 shows that the addition of GGBS also increases the
compressive strength. The higher reactivity of GGBS accommodates
faster dissolution enabling the early formation of the geopolymer
network and subsequently high early strength, irrespective of alkaline
(Lee et al., 2016; Mohamed, Al Khattab, and Al Hawat 123AD). On the
other hand, with the diminution of the Na2SiO3/GGBS ratio, the strength
improves, as shown in Fig. 6. Based on the mechanical properties, the
RM50-AS allows the highest incorporation of RM and ensures high early,
as well as 28-day compressive strength. Therefore, RM50-AS is the
optimized mixture for developing high-strength geopolymers consid­
ering the maximum utilization of red mud.
The molar ratios of Na/Al and Si/Na play a vital role in the geo­
polymerization process as they denote the availability of Na+ ions for Fig. 5. Effect of initial molar ratio on compressive strength of red mud (RM)-
both dissolution and activation of aluminosilicates. After the based geopolymers.

Fig. 4. SEM micrographs of raw materials, (1) red mud (RM), and (2) ground granulated blast-furnace slag (GGBS).

4
U. Zakira et al. Journal of Cleaner Production 383 (2023) 135439

CaO/SiO2 was found to be 0.64 for high strength RM- and GGBS-based
geopolymer as shown in Fig. 8. In summary, both a proper balance be­
tween the molar ratios and the method of mixing are key factors for
geopolymerization. The strength and molar ratios listed in Table 3
(Method B) prove the interdependency between the ion ratios; the
mixing procedure in method B might cause a hindrance in the conden­
sation process.

3.3. Microstructural properties of geopolymer

After 28 days of curing at 60 ◦ C, SEM was used to investigate the


morphological characteristics of geopolymers derived from RM and
GGBS. The micrographs of the geopolymers synthesized by Method A
and Method B are shown in Figs. 9 and 10, respectively.
Alkaline activation dissolves aluminosilicates, leading to the for­
mation of tetrahedral-shaped polymers and three-dimensional chain
networks that link the polymers. Compared to the SEM micrographs of
the raw materials shown in Fig. 4, the SEM images in Figs. 9 and 10 are
distinctly indicative of new phase formation. Geopolymer gels were
Fig. 6. Effect of ground granulated blast-furnace slag (GGBS) and Na2SiO3/ identified in all the samples, with a few partially reacted zones in be­
GGBS ratio on compressive strength of red mud (RM)-based geopolymers.
tween the unreacted minerals. The tetrahedral polymer can be seen in
the SEM images of geopolymers RM10-AS, and RM70-AS, and the gel
aluminosilicates have been activated, the condensation process begins, network formations are visible in the geopolymer RM90-AS. The geo­
either between the Si4+ and Al3+ species, forming poly-sialate, or be­ polymers synthesized by Method A seemed to be denser, more con­
tween the Si4+ ions, followed by the Al3+ ions and resulting in a rigid 3D nected, and have fewer unreacted particles than those synthesized by
network of geopolymers. The former case occurs when the SiO2/Al2O3 Method B. The tetrahedral polymers, unreacted minerals, and partially
ratio is low, and the latter case occurs when the SiO2/Al2O3 ratio is high reacted particles are visible in Fig. 10. The morphological structure of
(Singh et al., 2018). An excess concentration of the Si4+ ion may hamper RM50-BS was more compact than that of RM70-BS and RM90-BS, which
the condensation process as happened in the RM90-AS sample in the was also reflected by the compressive strength. The CaO content
study. Thus, there is an optimal range of SiO2/Al2O3, SiO2/Na2O, increased upon calcination. The morphology of RM-90AS shows the gel
Na2O/Al2O3 that are suitable for geopolymer formation. Fig. 7 shows network initiation and is yet to be condensed same as the samples RM10-
that the optimum ratio of SiO2/Na2O is 3.25 based on the polynomial AS to RM70-AS, which was subsequently echoed in the strength prop­
interpretation. The compressive strength increases with a higher pro­ erties. The high CaO content in GGBS caused the formation of amor­
portion of GGBS and an increase in the SiO2/Na2O ratio, which agrees phous Ca–Al–Si gels during geopolymerization decreasing the
with previous studies focused on slag-based geopolymerization (Tan microstructural porosity and strengthening the geopolymers (Yip and
et al., 2020; Ali Shah et al., 2021). An increase of Si/Al ratio forms more Van Deventer, 2003; Khale and Chaudhary, 2007).
Si–O–Al bonds and results in an improvement in the compressive The mineral phase characterization of the alkali-activated RM- and
strength (E. H. Kim and Leslie Struble, 2012). In contrast, the optimum GGBS-based geopolymers was performed by X-ray powder diffraction
range for Na2O/Al2O3 and SiO2/Al2O3 is approximately 1.1–3.4 and (XRD) on powder samples that had been cured for 28 days. Figs. 11 and
4.41 to 9.4, respectively. Ca/Si and Si/Al influence the strength of the 12 depict the XRD patterns of geopolymers comprised of identical pro­
geopolymer significantly. In the current study, the optimum value of portions of the same raw materials that were prepared using Methods A

Fig. 7. Effect of initial SiO2/Na2O on 28-day compressive strength of red mud Fig. 8. Effect of initial CaO/SiO2 on 28-day compressive strength of red mud
(RM)-based geopolymers towards determining optimum molar ratios with (RM)-based geopolymers towards determining optimum molar ratios with
polynomial correlation. polynomial correlation.

5
U. Zakira et al. Journal of Cleaner Production 383 (2023) 135439

Fig. 9. SEM micrographs of red mud (RM)-based geopolymers prepared by Method A and cured for 28 days: (a) RM10-AS, (b) RM30-AS, (c) RM50-AS, (d) RM50-
ASF, (e) RM70-AS, (f) RM90-AS.

Fig. 10. SEM micrographs of red mud (RM)-based geopolymers prepared by Method B and cured for 28 days: (a՛) RM50-BS, (b՛) RM70-BS, (c՛) RM90-BS.

and B, respectively. The large hump between 25o to 35o (2θ) in the XRD has more calcium and alumina, is likely to form calcium aluminum oxide
pattern of the raw GGBS denotes the amorphous phase. In contrast, no hydrate and CSH after complete dissolution of pyrophyllite, and exhibits
broad hump is visible in the raw RM samples; the major phases in RM are enhanced mechanical strength. Fig. 12 shows that geopolymers devel­
hematite (Fe2O3) at 33◦ –37◦ (2θ), quartz (SiO2) at 22◦ (2θ), and calcite oped by Method B possess five distinct peaks: calcium aluminum oxide
(CaCO3) at 29◦ (2θ). There were distinct peaks in the geopolymers of hydrate at 10◦ –12◦ (2θ), katoite (Ca2.93Al1.97Si0.64O2.56(OH)9.44) at
both sets of samples, and calcium silicate hydrate (CSH) was the major 17◦ –18◦ (2θ), 32◦ –32.7◦ (2θ), and 44.4◦ –46◦ (2θ), pyrophyllite
reaction product in both. Previous studies found that calcium-rich (Al4Si8O20(OH)4)0.33) at 28.5◦ –29.3◦ (2θ) and orientite (Ca2M­
feedstock, e.g., blast furnace slag, produces a calcium silicate hydrate n3O10(OH)4. The geopolymers prepared with Method A did not form
(CSH) gel when activated with an alkaline solution (Khale and katoite; rather a pyrophyllite peak was found in both of the geopolymers
Chaudhary, 2007). Fig. 11 shows that when the proportion of RM is less made with Methods A and B. Katoite forms in a CaO–SiO2–Al2O3–H2O
than 50% (the ratio of RM to GGBS ratio is below 1:1), calcium system when the silica content is sufficiently low (Meller et al., 2005).
aluminum oxide hydrate appears and periclase (MgO) pyrophyllite This observation reveals that due to more favorable geopolymerization,
(Al4Si8O20(OH)4)0.33) and orientite (Ca2Mn3O10(OH)4 disappear. In more calcium, aluminum, and silicon-based phases were formed by the
earlier studies, Yip and Van Deventer 2003 found that various types of Method A process than by the Method B process. The impact of this
calcium silicate hydrate, calcium aluminate hydrate, calcium alumino­ phenomenon was proven by the higher compressive strength of the
silicate hydrate, and CSH are prevalent in high-weight percent GGBS geopolymers developed by Method A.
systems, and the calcium aluminum oxide hydrate (CA) phase enhances
the degree of polycondensation and leads to more micro-crosslinking
geopolymer networks that strengthen the geopolymers (Moudio et al., 3.4. Effect of silica fume (SF) on RM-GGBS-based geopolymers
2021). The current study agrees with their findings. Pyrophyllite leaches
out Al3+ and Si4+ in an alkaline medium to form tetrahedral oligomeric To assess the effects of silica fume on the microstructure and me­
silicate or aluminate (Panda et al., 2018), and periclase, a chanical strength of RM-GGBS geopolymers, 10% of the GGBS in RM50-
magnesium-rich phase, forms when there is an abundance of magnesium AS was replaced with silica fume, which resulted in initial molar ratios
in the geopolymer ingredients. Thus, a matrix with less than 50% RM similar to those of the RM50-AS sample. The mix proportion was 50%
RM, 40% GGBS, and 10% silica fume, as shown in Table 1. The 28-day

6
U. Zakira et al. Journal of Cleaner Production 383 (2023) 135439

Fig. 11. XRD patterns of red mud (RM)-based geopolymers developed by Method A: CA, G, C, H, S, O, P, CSH, Py, and Pe denote peaks of calcium aluminum oxide
hydrate, gaultite, calcite, hematite, tridymite, orientite, potassium oxide nitrite, calcium silicate hydrate, pyrophyllite, and periclase, respectively.

Fig. 12. XRD patterns of red mud (RM)-based geopolymers developed by Method B: CA, G, C, H, S, O, P, CSH, Py, Pe, and K denote peaks of calcium aluminum oxide
hydrate, gaultite, calcite, hematite, tridymite, orientite, potassium oxide nitrite, calcium silicate hydrate, pyrophyllite, and periclase, katoite, respectively.

compressive strength of the sample with silica fume (Table 3), was highly dense microstructure of RM50-ASF. The X-ray powder diffraction
equivalent to that of the RM10-AS and RM50-AS samples. The early (XRD) analysis depicted in Figs. 11 and 13 illustrate that the mineral
strength of the RM50-ASF geopolymer was 10% greater than the geo­ phases of RM50-ASF are calcium aluminum oxide hydrate, gaultite
polymer without silica fume. The SEM images shown in Fig. 9 show the (Na4(Zn2Si7O18)(H2O)5), orientate (Ca2Mn3O10(OH)), hematite,

7
U. Zakira et al. Journal of Cleaner Production 383 (2023) 135439

Fig. 13. Comparison of XRD patterns of geopolymers with 50% red mud (RM) with and without silica fume (SF): CA, G, C, H, S, O, P, CSH, Py, Pe, and K denote peaks
of calcium aluminum oxide hydrate, gaultite, calcite, hematite, tridymite, orientite, potassium oxide nitrite, calcium silicate hydrate, pyrophyllite, periclase, and
katoite, respectively.

tridymite, potassium oxide nitrite (K4O(NO2)2), and calcium silicate framework is full of challenges. Second, the cost, performance, and
hydrate. No pyrophyllite, periclase, or katoite were present in the RM50- impacts of solid wastes can be localized or site-specific whereas the
ASF geopolymer, which is identical to the geopolymers prepared with current LCA typically adopts general values which simplified them.
10%RM (RM10-AS) and 30%RM (RM-30). The silica present in RM is Finally, the impact of the landfill and accumulation of solid wastes on
typically non-reactive (Singh et al., 2018), yet an additional source of the environment and how they deteriorate the natural environment or
active silica needs to be incorporated to synthesize high-strength geo­ assets are barely understood. It is extremely hard to track the trace of
polymers. Silica fume increases the reactivity of alkali reagents, favor­ hazardous substances released from the landfilled solid wastes.
ably Na+ ions in the alkali reagents, and surges the reactivity of the slag Furthermore, it is challenging to realize an accurate calculation due to
or GGBS by enhancing the reaction of Al3+ in GGBS (Lee et al., 2016). In the lack of appropriate correlation between the laboratory testing data
the current study, as shown in Fig. 13, the high reactivity of silica fume and field performance data.
and the presence of CaO in GGBS enabled the formation of calcium
aluminum oxide hydrate and C–S–H phases after the complete dissolu­ 5. Conclusion
tion of pyrophyllite, resulting in a high-strength geopolymer that was
equivalent to the geopolymers with a lower proportion of RM, irre­ The potential use of industrial by-products and wastes in the devel­
spective of the optimum range of the initial molar ratio. opment of high-strength geopolymers was investigated in this study.
Two mixing methods were performed to activate the aluminosilicate
4. Future recommendation precursors, and silica fume was introduced as a highly efficient additive
in the geopolymerization of RM and the GGBS system. This study can
Although the mechanical properties of the alkali-activated RM show greatly contribute to the advantageous reuse of huge RM. The study
a promising potential if combined with appropriate solid wastes and presents the mechanical and microstructural experimental testing re­
activators, the life cycle assessment (LCA) of these RM-based cementi­ sults of RM- and GGBS-based geopolymers and drws the following
tious materials is imperative before the real practical applications. conclusions.
Previous studies have demonstrated that sustainability takes three di­
mensions in the economic, social, and environmental domains (Ramos (1) The RM10-AS geopolymer, which was synthesized with 10% RM
and Abel, 2022); as a result, it is critical to assemble the knowledge and and 90% GGBS, had the highest level of compressive strength:
bridge the research gaps that are relevant to the durability of solid waste 68.8 MPa within 3 days by curing at 60 ◦ C ambient temperature
recycling from each of these three pillars to realize the alkali-activated for an initial molar ratio of SiO2/Al2O3 = 4.41, SiO2/Na2O =
RM-based products and strategies. The potential sources resulting in 4.02, Na2O/Al2O3 = 1.1, and CaO/SiO2 = 0.78. However, RM50-
such complexities in the LCA study may include, but are not limited to, AS is the optimum geopolymer, based on the utilization of 50%
the issues discussed as follows. First, the indirect implications in the RM and achievement of a high strength of 65.7 MPa.
environmental footprint and cost, which raises the need to define the (2) Based on the polynomial fit, the optimum molar ratio of SiO2/
boundaries and time scale of the analysis domain to select the appro­ Na2O and CaO/SiO2 is 3.25 and 0.64, respectively. The optimum
priate temporal and spatial resolution for the LCA study. How to set up range for Na2O/Al2O3 and SiO2/Al2O3 was found to be 1.1 to 3.4
the boundary of research to cover the major considerations in the LCA and 4.41 to 9.4, respectively.

8
U. Zakira et al. Journal of Cleaner Production 383 (2023) 135439

(3) The geopolymer samples prepared with premixing with NaOH metakaolin-based geopolymers. Colloids Surf. A Physicochem. Eng. Asp. 292 (1),
8–20. https://doi.org/10.1016/j.colsurfa.2006.05.044.
solution (Method A) showed better mechanical and microstruc­
GB/T 50107-2010, 2010. Standard for Evaluation of Concrete Compressive Strength
tural performances than Method B where the NaOH and Na2SiO3 (Chinese Standard). China Architecture Industry Press, Beijing, China.
solution were introduced altogether. The geopolymers synthe­ Gräfe, M., Power, G., Klauber, C., 2011. Bauxite residue issues: III. Alkalinity and
sized by Method A were denser, more connected, and had fewer associated chemistry. Hydrometallurgy 108, 60–79. https://doi.org/10.1016/j.
hydromet.2011.02.004.
unreacted particles than those prepared by Method B. The SEM Hairi, Siti Noor Md., Jameson, Guy N.L., Rogers, Joanne J., MacKenzie, Kenneth J.D.,
images showed that lower weight percent of RM in the raw in­ 2015. Synthesis and properties of inorganic polymers (geopolymers) derived from
gredients resulted in geopolymer matrices that were more bayer process residue (red mud) and bauxite. J. Mater. Sci. 50 (23), 7713–7724.
https://doi.org/10.1007/S10853-015-9338-9/FIGURES/10.
condensed and compact, which also justified the strength results. Hanif, Asad, Kim, Yongjae, Lu, Zeyu, Park, Cheolwoo, 2017. Early-age behavior of
(4) With an RM to GGBS ratio below 1:1, XRD analysis showed that recycled aggregate concrete under steam curing regime. J. Clean. Prod. 152,
the hydrated calcium-aluminum phase appeared and periclase 103–114. https://doi.org/10.1016/j.jclepro.2017.03.107.
He, Jian, Yuxin, Jie, Zhang, Jianhong, Yu, Yuzhen, Zhang, Guoping, 2013. Synthesis and
disappeared, which enhanced the mechanical strength. The characterization of red mud and rice husk ash-based geopolymer composites.
higher strength of the geopolymers developed by the Method A Cement Concr. Compos. 37 (1), 108–118. https://doi.org/10.1016/j.
was associated with enhanced Al3+-based aluminosilicate phase cemconcomp.2012.11.010.
Hertel, Tobias, Pontikes, Yiannis, 2020. Geopolymers, inorganic polymers, alkali-
formation; therefore, Method A was proven to be more efficient activated materials and hybrid binders from bauxite residue (red mud) – putting
than Method B things in perspective. J. Clean. Prod. 258 (June), 120610 https://doi.org/10.1016/J.
(5) Silica fume (SF) was proven to be an effective replacement for JCLEPRO.2020.120610.
Hu, Wei, Nie, Qingke, Huang, Baoshan, Xiang, Shu, He, Qiang, 2018a. Mechanical and
GGBS, due to its increased reactivity that resulted in higher early
microstructural characterization of geopolymers derived from red mud and fly ashes.
compressive strength. A high weight percent of RM with 10 wt J. Clean. Prod. 186 (June), 799–806. https://doi.org/10.1016/J.
percent of SF was utilized and resulted in high strength of 67.3 JCLEPRO.2018.03.086.
MPa. Hu, Wei, Nie, Qingke, Huang, Baoshan, Su, Aiwu, Du, Yong, Xiang, Shu, He, Qiang,
2018b. Mechanical property and microstructure characteristics of geopolymer
stabilized aggregate base. Construct. Build. Mater. 191 (December), 1120–1127.
CRediT authorship contribution statement https://doi.org/10.1016/J.CONBUILDMAT.2018.10.081.
Hu, Wei, Nie, Qingke, Huang, Baoshan, Xiang, Shu, 2019. Investigation of the strength
development of cast-in-place geopolymer piles with heating systems. J. Clean. Prod.
Umme Zakira: Methodology, Data curation, Data interpretation, 215, 1481–1489. https://doi.org/10.1016/j.jclepro.2019.01.155.
Writing – original draft, reviewing, and editing. Kai Zheng: Methodol­ Jiang, Xi, Xiao, Rui, Zhang, Miaomiao, Hu, Wei, Bai, Yun, Huang, Baoshan, 2020a.
ogy, Data curation, Data interpretation. Ning Xie: Conceptualization, A laboratory investigation of steel to fly ash-based geopolymer paste bonding
behavior after exposure to elevated temperatures. Construct. Build. Mater. 254
Visualization, Supervision. Bjorn Birgisson: Conceptualization, Visu­ (September), 119267 https://doi.org/10.1016/J.CONBUILDMAT.2020.119267.
alization, Supervision. Jiang, Xi, Zhang, Yiyuan, Xiao, Rui, Polaczyk, Pawel, Zhang, Miaomiao, Hu, Wei,
Bai, Yun, Huang, Baoshan, 2020b. A comparative study on geopolymers synthesized
by different classes of fly ash after exposure to elevated temperatures. J. Clean. Prod.
Declaration of competing interest 270 (October), 122500 https://doi.org/10.1016/J.JCLEPRO.2020.122500.
Khale, Divya, Chaudhary, Rubina, 2007. Mechanism of geopolymerization and factors
The authors declare that they have no known competing financial influencing its development: a review. J. Mater. Sci. 42 (3), 729–746. https://doi.
org/10.1007/s10853-006-0401-4.
interests or personal relationships that could have appeared to influence
Kim, Eric Heysung, Leslie Struble, 2012. Understanding Effects of Silicon/Aluminum
the work reported in this paper. Ratio and Calcium Hydroxide on Chemical Composition, Nanostructure and
Compressive Strength for Metakaolin Geopolymers. https://www.ideals.illinois.ed
Data availability u/handle/2142/34257.
Kim, Yongjae, Hanif, Asad, Usman, Muhammad, Junaid Munir, Muhammad, Saleem
Kazmi, Syed Minhaj, Kim, Samsoo, 2018. Slag waste incorporation in high early
Data will be made available on request. strength concrete as cement replacement: environmental impact and influence on
hydration & durability attributes. J. Clean. Prod. 172, 3056–3065. https://doi.org/
10.1016/j.jclepro.2017.11.105.
Acknowledgments Kumar, Aman, Jothi Saravanan, T., Bisht, Kunal, Syed Ahmed Kabeer, K.I., 2021.
A review on the utilization of red mud for the production of geopolymer and alkali
We would like to thank Jinmei Shi and Yuanchao Chen for their help activated concrete. Construct. Build. Mater. 302 (October), 124170 https://doi.org/
10.1016/J.CONBUILDMAT.2021.124170.
during the lab test. Lee, N.K., An, G.H., Koh, K.T., Ryu, G.S., 2016. Improved reactivity of fly ash-slag
geopolymer bythe addition of silica Fume.Pdf. Adv. Mater. Sci. Eng. 11 https://doi.
References org/10.1155/2016/2192053, 2016.
Li, Shucai, Zhang, Jian, Li, Zhaofeng, Gao, Yifan, Liu, Chao, Mo, Bing Hui, He, Zhu, et al.,
2018. Recycling and utilization assessment of waste fired clay bricks (grog) with
ACI 363R, 2010. Report on High-Strength Concrete. Farmington Hills, MI, USA. www.co
granulated blast-furnace slag for geopolymer production. Construct. Build. Mater.
ncrete.org/committees/errata.asp.
267 (1–2), 317–325. https://doi.org/10.1016/j.conbuildmat.2020.120942.
Agrawal, Varsha, Paulose, Rini, Arya, Rahul, Rajak, Gaurav, Giri, Abhishek,
Li, Yuancheng, Min, Xiaobo, Ke, Yong, Liu, Degang, Tang, Chongjian, 2019. Preparation
Bijanu, Abhijit, Sanghi, Sunil K., et al., 2022. Green conversion of hazardous red
of red mud-based geopolymer materials from MSWI fly ash and red mud by
mud into diagnostic X-ray shielding tiles. J. Hazard Mater. 424 (February), 127507
mechanical activation. Waste Manag. 83 (January), 202–208. https://doi.org/
https://doi.org/10.1016/J.JHAZMAT.2021.127507.
10.1016/J.WASMAN.2018.11.019.
Andrew, Robbie M., 2018. Global CO2 emissions from cement production. Earth Syst.
Li, Shucai, Zhang, Jian, Li, Zhaofeng, Gao, Yifan, Liu, Chao, 2021. Feasibility study of red
Sci. Data 10, 195–217. https://doi.org/10.5194/essd-10-195-2018.
mud-blast furnace slag based geopolymeric grouting material: effect of
ASTM C188-16, 2009. Standard test method for density of hydraulic cement. ASTM
superplasticizers. Construct. Build. Mater. 267, 120910 https://doi.org/10.1016/j.
International 95 (C), 1–3. https://doi.org/10.1520/C0188-16.2.
conbuildmat.2020.120910.
ASTM C39/C39M, 2003. Standard Test Method for Compressive Strength of Cylindrical
Liguori, Barbara, Capasso, Ilaria, De Pertis, Marco, Ferone, Claudio, Cioffi, Raffaele,
Concrete Specimens 1. https://doi.org/10.1520/C0039. ASTM Standard Book i
2017. Geopolymerization Ability of Natural and Secondary Raw Materials by
(March): 1–5.
Solubility Test in Alkaline Media. https://doi.org/10.3390/environments4030056.
Capasso, Ilaria, Liguori, Barbara, Ferone, Claudio, Caputo, Domenico, Cioffi, Raffaele,
Liu, Wanchao, Chen, Xiangqing, Li, Wangxing, Yu, Yanfen, Yan, Kun, 2014.
2020. Strategies for the Valorization of Soil Waste by Geopolymer Production: an
Environmental assessment , management and utilization of red mud in China.
Overview. https://doi.org/10.1016/j.jclepro.2020.125646.
J. Clean. Prod. 84, 606–610. https://doi.org/10.1016/j.jclepro.2014.06.080.
Chindaprasirt, P., Chareerat, T., Hatanaka, S., Cao, T., 2011. High-strength geopolymer
Mclellan, Benjamin C., Williams, Ross P., Lay, Janine, Van Riessen, Arie, Corder, Glen D.,
using fine high-calcium fly ash. J. Mater. Civ. Eng. 23 (3), 264–270. https://doi.org/
2011. Costs and Carbon Emissions for Geopolymer Pastes in Comparison to Ordinary
10.1061/(asce)mt.1943-5533.0000161.
Portland Cement. https://doi.org/10.1016/j.jclepro.2011.02.010.
Davidovits, Joseph, Davidovics, Michel, Davidovits, Nicolas, 1991. Process for Obtaining
Meller, Nicola, Hall, Christopher, Phipps, Jonathan S., 2005. A new phase diagram for
a Geopolymeric Alumino Silicate and Products Thus Obtained (September).
the CaOAl2O3SiO 2H2O hydroceramic system at 200 ◦ c. Mater. Res. Bull. 40 (5),
Davidovits, J., Davidovics, M., Davidovits, N., 1994. Process for obtaining a
715–723. https://doi.org/10.1016/j.materresbull.2005.03.001.
geopolymeric alumino-silicate. US Patent 5 (595), 342 issued 1994.
Mohamed, Osama A, Rania Al Khattab, and Waddah Al Hawat. 123AD. “Effect of
Duxson, P., Mallicoat, S.W., Lukey, G.C., Kriven, W.M., van Deventer, J.S.J., 2007. The
Relative GGBS/Fly Contents and Alkaline Solution Concentration on Compressive
effect of alkali and Si/Al ratio on the development of mechanical properties of

9
U. Zakira et al. Journal of Cleaner Production 383 (2023) 135439

Strength Development of Geopolymer Mortars Subjected to Sulfuric Acid.” htt Vigneshwaran, S., Uthayakumar, M., Arumugaprabu, V., 2019. Development and
ps://doi.org/10.1038/s41598-022-09682-z. sustainability of industrial waste-based red mud hybrid composites. J. Clean. Prod.
Morsy, M.S., Alsayed, S.H., Al-Salloum, Y., Almusallam, T., 2014. Effect of sodium 230, 862–868. https://doi.org/10.1016/j.jclepro.2019.05.131.
silicate to sodium hydroxide ratios on strength and microstructure of fly ash Vigneshwaran, S., Uthayakumar, M., Arumugaprabu, V., 2020. Potential use of industrial
geopolymer binder. Arabian J. Sci. Eng. 39 (6), 4333–4339. https://doi.org/ waste-red mud in developing hybrid composites : a waste management approach.
10.1007/S13369-014-1093-8, 2014 39:6. J. Clean. Prod. 276, 124278 https://doi.org/10.1016/j.jclepro.2020.124278.
Moudio, A.M.N., Tchakouté, H.K., Ngnintedem, D.L.V., Andreola, F., Kamseu, E., Wang, Shaohan, Jin, Huixin, Deng, Yong, Xiao, Yuandan, 2021. Comprehensive
Nanseu-Njiki, C.P., Leonelli, C., Rüscher, C.H., 2021. (No title). Mater. Chem. Phys. utilization status of red mud in China : a critical review. J. Clean. Prod. 289, 125136
264, 124459 https://doi.org/10.1016/j.matchemphys.2021.124459. https://doi.org/10.1016/j.jclepro.2020.125136.
Murayama, Norihiro, Yamamoto, Hideki, Shibata, Junji, 2002. Mechanism of zeolite Wang, Yanhai, Xiao, Rui, Hu, Wei, Jiang, Xi, Zhang, Xiao, Huang, Baoshan, 2021. Effect
Synthesis from coal fly ash by alkali hydrothermal reaction. Int. J. Miner. Process. 64 of granulated phosphorus slag on physical, mechanical and microstructural
(1), 1–17. https://doi.org/10.1016/S0301-7516(01)00046-1. characteristics of class F fly ash based geopolymer. Construct. Build. Mater. 291
National Minerals Information Center, Usgs. 2020. Mineral Commodity Summaries. (July), 123287 https://doi.org/10.1016/J.CONBUILDMAT.2021.123287.
Mineral Commodity Summaries. https://doi.org/10.3133/MCS2020. Xiao, Rui, Ma, Yuetan, Jiang, Xi, Zhang, Miaomiao, Zhang, Yiyuan, Wang, Yanhai,
Neupane, Kamal, 2018. High-strength geopolymer concrete- properties, advantages and Huang, Baoshan, He, Qiang, 2020. Strength, microstructure, efflorescence behavior
challenges. Adv. Mater. 7 (2), 15. https://doi.org/10.11648/j.am.20180702.11. and environmental impacts of waste glass geopolymers cured at ambient
Nie, Qingke, Hu, Wei, Tao, Ai, Huang, Baoshan, Xiang, Shu, He, Qiang, 2016. Strength temperature. J. Clean. Prod. 252 (April), 119610 https://doi.org/10.1016/J.
properties of geopolymers derived from original and desulfurized red mud cured at JCLEPRO.2019.119610.
ambient temperature. Construct. Build. Mater. 125, 905–911. https://doi.org/ Xiao, Rui, Zhang, Yiyuan, Jiang, Xi, Polaczyk, Pawel, Ma, Yuetan, Huang, Baoshan, 2021.
10.1016/j.conbuildmat.2016.08.144. Alkali-activated slag supplemented with waste glass powder: laboratory
Panda, Laxmipriya, Rath, Swagat S., Srinivas Rao, Danda, Nayak, Binod B., characterization, thermodynamic modelling and sustainability analysis. J. Clean.
Das, Bisweswar, Misra, Pramila K., 2018. Thorough understanding of the kinetics Prod. 286 (March), 125554 https://doi.org/10.1016/J.JCLEPRO.2020.125554.
and mechanism of heavy metal adsorption onto a pyrophyllite mine waste based Xu, Hua, Van Deventer, Jannie S.J., 2003. Effect of source materials on
geopolymer. J. Mol. Liq. 263 (June), 428–441. https://doi.org/10.1016/j. geopolymerization. Ind. Eng. Chem. Res. 42 (8), 1698–1706. https://doi.org/
molliq.2018.05.016. 10.1021/ie0206958.
Paramguru, R.K., Rath, P.C., Misra, V.N., 2004. Trends in the red mud utilization-A Ye, Nan, Yang, Jiakuan, Ke, Xinyuan, Zhu, Jing, Li, Yalin, Cheng, Xiang, Wang, Huabin,
review. Miner. Process. Extr. Metall. Rev. 26, 1. https://doi.org/10.1080/ Li, Lei, Xiao, Bo, 2014. Synthesis and characterization of geopolymer from bayer red
08827500490477603, 1. mud with thermal pretreatment. J. Am. Ceram. Soc. 97 (5), 1652–1660. https://doi.
Ramos, Ana, Abel, Rouboa, 2022. Life cycle thinking of plasma gasification as a waste-to- org/10.1111/JACE.12840.
energy tool: review on environmental, economic and social aspects. Renew. Sustain. Ye, Nan, Yang, Jiakuan, Liang, Sha, Hu, Yong, Hu, Jingping, Xiao, Bo, Huang, Qifei,
Energy Rev. 153 (January), 111762 https://doi.org/10.1016/J.RSER.2021.111762. 2016. Synthesis and strength optimization of one-Part Geopolymer based on red
Rowles, Matthew, O’Connor, Brian, 2003. Chemical optimisation of the compressive mud. Construct. Build. Mater. 111, 317–325. https://doi.org/10.1016/j.
strength of aluminosilicate geopolymers synthesised by sodium silicate activation of conbuildmat.2016.02.099.
metakaolinite. J. Mater. Chem. 13 (5), 1161–1165. https://doi.org/10.1039/ Yip, C.K., Van Deventer, J.S.J., 2003. Microanalysis of calcium silicate hydrate gel
B212629J. formed within a geopolymeric binder. J. Mater. Sci. 38, 3851–3860.
Shah, Ali, Farasat, Syed, Chen, Bing, Ahmad, Muhammad Riaz, Aminul Haque, M., 2021. Zhang, Yiyuan, Xiao, Rui, Jiang, Xi, Li, Wenkai, Zhu, Xingyi, Huang, Baoshan, 2020.
Development of cleaner one-Part Geopolymer from lithium slag. J. Clean. Prod. 291 Effect of particle size and curing temperature on mechanical and microstructural
(April), 125241 https://doi.org/10.1016/J.JCLEPRO.2020.125241. properties of waste glass-slag-based and waste glass-fly ash-based geopolymers.
Singh, Smita, Aswath, M.U., Ranganath, R.V., 2018. Effect of mechanical activation of J. Clean. Prod. 273 (November), 122970 https://doi.org/10.1016/J.
red mud on the strength of geopolymer binder. Construct. Build. Mater. 177, JCLEPRO.2020.122970.
91–101. https://doi.org/10.1016/j.conbuildmat.2018.05.096. Zinoveev, Dmitry, Pasechnik, Liliya, Fedotov, Mikhail, Dyubanov, Valery,
Tan, Jiawei, Cai, Jingming, Li, Xiaopeng, Pan, Jinlong, Li, Jiabin, 2020. Development of Grudinsky, Pavel, Alpatov, Andrey, 2021. Extraction of valuable elements from red
eco-friendly geopolymers with ground mixed recycled aggregates and slag. J. Clean. mud with a focus on using liquid media—a review. Recycling 6, 38. https://doi.org/
Prod. 256 (May), 120369 https://doi.org/10.1016/J.JCLEPRO.2020.120369. 10.3390/RECYCLING6020038, 2021, Page 38 6 (2).

10

You might also like