Next Article in Journal
User-Oriented Sustainable Renewal of Peri-Urban Heritage Towns: A Case Study of Nanquan Street, Wuxi, China
Previous Article in Journal
Bioprospecting Native Oleaginous Microalgae for Wastewater Nutrient Remediation and Lipid Production: An Environmentally Sustainable Approach
Previous Article in Special Issue
Assessment of the Efficiency of Mechanical Grinding and Calcination Processes for Construction and Demolition Waste as Binder Replacement in Cement Pastes: Mechanical Properties Evaluation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Fly Ash- and Slag-Based Geopolymer Cements

by
Taofiq O. Mohammed
,
Aman Ul Haq
,
Mohammad Zunaied Bin Harun
and
Ebenezer O. Fanijo
*
School of Building Construction, Georgia Institute of Technology, Atlanta, GA 30332, USA
*
Author to whom correspondence should be addressed.
Sustainability 2025, 17(24), 11167; https://doi.org/10.3390/su172411167
Submission received: 29 September 2025 / Revised: 28 November 2025 / Accepted: 28 November 2025 / Published: 12 December 2025

Abstract

This review study promotes the sustainability of civil infrastructure by advancing the materials science of alternative cementitious materials. Supported by extensive global research and industrial trials, geopolymer cement has emerged as a promising approach to reducing the ecological impact of ordinary Portland cement (OPC) due to its superior engineering properties and eco-friendly benefits from industrial waste utilization. Geopolymers are inorganic polymers formed by the polymerization of aluminosilicate precursors, such as fly ash (FA), slag, and metakaolin, in the presence of alkaline activating solutions. This work integrates findings across multiple domains, including precursor chemistry, microstructural evolution, mechanical and durability performance, and sustainability metrics like carbon footprint and energy consumption. A key contribution of this review is the comparative evaluation of FA-based and slag-based GPC systems against OPC concrete, emphasizing the factors influencing their mechanical and durability properties, while also distinguishing differences in environmental impact, microstructural development, and overall performance. The findings highlight that slag-based systems generally exhibit lower environmental impacts, especially in energy demand and emissions, while regional differences in precursor availability constrain how widely the LCA and economic results can be applied. Building on previous reviews that have considered these topics, this study jointly examines technical performance and sustainability indicators and identifies regional variations that influence feasibility. The synthesis provides a balanced, evidence-based assessment of the potential and limitations of GPC as a lower-carbon alternative to OPC, supporting efforts to reduce the climate impact of future concrete construction.

1. Introduction

Presently, concrete is the most consumed material in the world after water, and it is undoubtedly the most essential building and construction material [1,2]. Ordinary Portland Cement (OPC) concrete is utilized in construction twice as much as all other building materials combined, and this demand is evident in its global production, which exceeds 30 billion tons each year [1,2,3,4]. The demand for concrete will continue to rise with the continuous increase in the rate of industrialization and urbanization, resulting in a rise in the need for infrastructure [5,6]. Concrete is composed of cement (binder), sand (fine aggregate), crushed stone or gravel (coarse aggregate), water, as well as sometimes supplementary cementitious material (SCMs) and chemical admixtures [1,5]. Concrete exhibits outstanding engineering properties, and when it is properly designed, the material is versatile, durable, and economical, and provides certain environmental benefits. Its most remarkable characteristic is that, unlike any other available material, it is an engineered product that can be tailored to meet nearly any conceivable set of performance standards. This adaptability makes it a crucial component in modern construction and infrastructure [1]. However, due to its enormous consumption, concrete and cement are also significant contributors to environmental problems and climate change, responsible for more than 8–10% of global anthropogenic CO2 emissions [7,8]. This high share stems mainly from the production of clinker, the key intermediate in cement manufacture, which releases approximately 0.85 tons of CO2 per ton of clinker produced through both limestone calcination and fuel combustion. With an average clinker-to-cement ratio of 0.72–0.75, global cement production exceeding 4 BMT annually emits about 2 BMT of CO2, as shown in Figure 1 [9,10,11], equivalent to roughly 250–300 kg of CO2 per m3 of conventional concrete.
The manufacturing process of cement is highly energy-intensive, involving the acquisition, transportation, grinding, mixing, and heating of raw materials, such as limestone, clay, and shale, in a cement kiln at temperatures up to 1500 °C to produce clinker. The clinker is then crushed and blended to produce cement [12,13]. For instance, the CO2 emissions result mainly from the decomposition of limestone (main reaction: CaCO3 → CaO + CO2) to produce CaO (the main content of clinker) and CO2, and well as from the burning of fossil fuels during cement production to provide the thermal energy required (heated to around 1400–1500 °C) for calcination to occur. Moreover, cement and concrete production require overexploitation of quarries, which also has a significant impact on landscape pollution. Alongside environmental motivations, evolving policy frameworks and corporate sustainability commitments are further driving research into alternative cementitious systems. To contribute to sustainable development, immediate action is needed to minimize CO2 emissions from cement manufacturing. Minimizing cement and concrete consumption is also paramount to achieving the net-zero emission scenario outlined in the International Energy Agency (IEA) report, which emphasizes decarbonization of the built environment as a critical global priority.
Consequently, cement scientists and manufacturers have begun to emphasize both improving the sustainability of OPC production and exploring alternative binders as complementary pathways toward achieving net-zero targets [14]. Some of the elucidated alternative binders to OPC technology include calcium aluminate cements, magnesium cements (CACs), calcium sulfoaluminate cements (CSA), supersulfated cements, the magnesium phosphate system, and geopolymer cement [15,16,17,18]. Compared to other binder technologies, geopolymer cement has a longer in-service track record and offers enormous potential for a sustainable future and climate change mitigation for structural engineering applications [19], supported by the extensive body of fundamental research relating gel chemistry and microstructure to their performance [20]. Aside from being a perfect replacement candidate, geopolymer cements are more durable, with excellent resistance to acid and sulfate attack, fire resistance, low thermal conductivity, extended service life, and suitability for sustainability, and they offer an eco-friendly disposal option for waste materials without the usual practice of landfill disposal.
Joseph Davidovits, a pioneering researcher in geopolymer concrete (GPC), developed this inorganic material in the 1970s using industrial waste products [21]. This process, known as geopolymerization, involves chemically activating aluminosilicate-rich industrial waste with an alkaline solution to form a durable and sustainable binder [22,23]. The reactive aluminosilicate precursors, such as fly ash (FA) [24,25], metallurgical (including blast furnace) slag [25,26], metakaolin (MK) [27,28], calcine clays, and/or volcanic ash [29], are commonly used in geopolymerization. The utilization of GPC as an alternative to OPC concrete has been reported to reduce CO2 emissions by up to 80% [30,31,32]. Since their development, there has been a significant increase in their application as an alternative material in concrete structures. For example, the largest GPC project in the world is Brisbane West Well Camp Airport (BWWA), Australia, where 70,000 tons of GPC were used for the pavement and floor panels, making it the world’s first greenfield public airport [33].
However, these conventional precursors have shortcomings associated with them, and these shortcomings will continue to limit their usage for the foreseeable future. For instance, FA, a byproduct of coal combustion that accumulates at the top of boilers [34,35,36], is experiencing a decline in availability due to the decreasing reliance on coal-fired power generation, which has led to a reduction in coal combustion residuals production [37]. Despite the reduction in the availability of FA, demand continues to rise, thereby leading to a market imbalance and a continuous rise in cost [37]. Similarly, slag, which is a byproduct of the steel-making industry [38,39,40], is also experiencing a recent reduction in its production due to the priority given to the use of recycled steel for steel production over the traditional production of steel [41,42,43,44]. As a result, its decreasing availability has driven up costs, further limiting its use. Meanwhile, MK, which is produced through the thermal activation of kaolin clay, has been reported to have high strength and durability compared to conventional concrete [45,46,47]. However, despite its excellent mechanical performance, MK-based GPC is uneconomical when compared to its prototype [44,48]. The excessive cost of MK is attributed to the low rate of production, due to the limited number of manufacturing plants [44].
The challenges associated with conventional precursors in GPC demand the urgent development and adoption of new SCMs that can replace traditional precursors while enabling the production of net-zero GPC to align with the 2050 carbon neutrality targets. Over the past three decades, numerous review articles have examined the mechanisms, physicochemical characteristics, strength, and material properties of GPC. However, despite the increasing emphasis on sustainable construction materials over the past decade and the push toward 2050 carbon neutrality goals, there remains a significant gap in comprehensive reviews that analyze the sustainability of different GPC systems and provide recommendations for achieving net-zero emissions.

Research Significance and Contribution

Several reviews have advanced the understanding of GPC systems but remain limited in scope. Recently, Patil et al. (2023) provided a broad overview of SCMs, such as FA, slag, MK, and rice husk ash, and their influence on mechanical properties, but did not extend their discussion to sustainability indicators, such as CO2 footprint or energy demand [49]. Venkatesan et al. (2024) offered a comprehensive analysis of Class F FA, alkaline activator molarity, and curing conditions on compressive strength, yet their review was narrowly restricted to siliceous fly ash-based systems, neglected durability properties and excluding other precursors and environmental dimensions [50]. El Alouani et al. (2024) [51] surveyed geopolymer synthesis, characterization, and their applications. Although their research was multidisciplinary, it did not address slag-based geopolymers or the role of GPC in advancing sustainability toward net-zero emissions [51]. Zhang (2024) concentrated on the durability performance of GPC, including acid, chloride, sulfate, and freeze–thaw resistance, and, while providing a critical durability perspective, did not consider recent development in life-cycle emissions or cost implications [52]. Similarly, Magotra and Jee (2024) highlighted the durability and microstructural features of FA-based GPC and presented its potential for reducing carbon emissions by up to 80%, but their scope was limited to FA systems, overlooking slag and other precursors, as well as regional variability in material availability and economic performance [53].
In contrast to earlier reviews that primarily focused on isolated aspects such as mechanical performance or chemical mechanisms, this study integrates findings across multiple domains, including precursor chemistry, microstructural evolution, mechanical and durability performance, and sustainability metrics like carbon footprint and energy consumption. A key contribution of this review is the comparative evaluation of FA-based and slag-based GPC systems against OPC concrete, emphasizing the factors influencing their mechanical and durability properties, while also distinguishing differences in environmental impact, microstructural development, and overall performance. Moreover, this review also explores how FA- and slag-based GPC are utilized from a global perspective and how their regional variability may affect the achievement of net-zero emissions. Alongside this, a structured methodology was adopted to provide a comprehensive understanding of the sustainable benefits of different GPC systems in comparison with conventional OPC concrete.
This review comprises nine sections, including the introduction (Section 1). Section 2 discusses the methodology employed, and Section 3 synthesizes advances in reaction chemistry and mechanisms, highlighting knowledge gaps. Section 4 examines physico-chemical properties and microstructure. Section 5 reviews workability and mechanical performance (compressive, tensile, flexural, and modulus of elasticity). Section 6 evaluates durability against acids, chlorides, sulfates, and corrosion, and their interrelations. Section 7 appraises methodological quality and limitations. Section 8 analyzes sustainability (energy use, CO2 emissions, economic viability, and regional cost variability), with comparisons of FA/slag and alternative precursors to conventional concrete. Section 9 concludes with key findings and future research directions toward net-zero geopolymer mixes.

2. Methodology

To systematically review the existing literature, this study followed the PRISMA protocol. The review process included the following steps: (1) defining the research question, (2) selecting authoritative databases and searching relevant literature, (3) screening for inclusion, (4) assessing the quality of the literature, and (5) review.

2.1. Research Questions

The following research questions were formulated to guide the systematic literature review process and set out the research objectives:
  • What are the key physio-chemical mechanisms and microstructural characteristics of fly ash- and slag-based geopolymer binders, and how do they differ from OPC systems?
  • How do the fresh and hardened properties of FA/slag-based geopolymers compare with OPC concrete?
  • How do fly ash- and slag-based geopolymer concretes perform under major durability treats (i.e., acid, chloride, sulfate attack, and steel corrosion) relative to OPC concrete?
  • From a cost and life-cycle sustainability perspective, under what regional and supply conditions can FA/slag-based geopolymer concretes be realistically adopted at a large scale as low-carbon alternatives to OPC?

2.2. Database Selection and Literature Search

In this study, the Web of Science database was selected due to its extensive and precise coverage of peer-reviewed publications. For the literature search, a total of 3106 publications were collated, of which 2728 were retained after removing review articles. To ensure quality, the review considered only publications in English with full-text availability. After screening based on titles and abstracts, the number was further reduced, and following assessment against the eligibility criteria, the set was narrowed again. The process is shown in Figure 2.

2.3. Literature Screening and Assessment

During screening and assessment, publications focused on other composite classes or approaches (e.g., OPC concrete merely citing geopolymer articles, computational tools, modeling) were excluded; only studies relevant to our research questions were selected. Overall, articles were included only if they characterized the following key aspects:
  • Form of geopolymer materials (i.e., paste, mortar, and concrete);
  • Type of pre-cursor material (i.e., fly ash, slag, fly ash + slag);
  • Chemical and microstructural parameters based on XRF, XRD, FTIR, TGA, SEM, and TEM tool data;
  • Physical properties of constituent materials based on density, water absorption, porosity;
  • Fresh properties based on workability.
  • Hardened properties based on compressive strength, tensile strength, flexural strength, modulus of elasticity.
  • Durability properties based on resistance to acid, sulfate, chloride, and corrosion;
  • Environmental impact based on regional variation.
  • Cost of production based on regional variation.

3. Background and Mechanism of GPC

3.1. Background of GPC

The discovery of a new class of inorganic material, or GPC binder, cement, and concrete, has resulted in wide scientific interest and kaleidoscopic development of applications due to their high mechanical strength and up to 80% lower CO2 emission compared to OPC concrete [54]. GPCs are inorganic polymers formed by the polymerization of aluminosilicate source material in the presence of alkaline activating solutions (i.e., activators such as alkali metal silicates, hydroxides, and carbonates) [20,55]. Among the various precursors used for GPC system production, two main reaction products, depending on precursors’ composition, are formed [56,57]. On the one hand, precursors rich in Si + Ca (e.g., blast furnace slag) form calcium-aluminate-silicate-hydrate (C-A-S-H) as the main reaction production in a low-alkaline solution [58]. On the other hand, precursors rich in Si + Al (e.g., metakaolin and fly ash) with low Ca content, produced inorganic polymers of sodium-aluminate-silicate-hydrate (N-A-S-H) [59]. The resulting products govern the properties of the cement from various perspectives, including mechanical strength, durability performance in acidic media, resistance to chloride penetration, sulfate attacks, and seawater. Another important constituent of geopolymerization is the activating solution, which consists of alkalis and sometimes reactive silica, like sodium hydroxide, sodium silicate, and potassium hydroxide, required to initiate the reaction process [60].
Geopolymers are a family of alumino-silicate binders, which are the result of an inorganic polycondensation reaction, a process known as geopolymerization. This reaction yields an amorphous to semicrystalline three-dimensional tecto-aluminosilicate frameworks with the general empirical formula [61] of Mn{−(SiO2)z–AlO2−}n. wH2O, where M is the cation or positive ion, such as sodium (Na+) or potassium (K+), calcium (Ca2+) is required in the structure to balance the negative charge, n is a degree of polycondensation, and the subscript z defines the Si-to-Al molar ratio, which is a vital factor in determining the engineering properties of the resulting geopolymers. Such frameworks are regarded as poly(sialates), where sialate is an abbreviation for the silicon-oxo-aluminate building unit. The sialate network is a combination of silicate and aluminate ([SiO4]4− and [AlO4]5−) tetrahedra, cross-linked by shared oxygen atoms [62]. Thus, poly(sialates) are chain and ring polymers with Si4+ and Al3+ in IV-fold coordination with oxygen, ranging from amorphous to semicrystalline, and are inaugurated as geopolymers [61]. Figure 3 describes the three main classes of polymeric chains commonly observed.
While these structural classifications are well established, the geopolymerization process itself is highly sensitive to raw material variability. The balance between precursor composition, activator chemistry, and curing conditions largely determines the rate of dissolution and condensation, the type of gel that forms, and the resulting microstructure [63]. For instance, low-calcium fly ash systems typically exhibit slower reactivity at ambient curing and often require elevated temperatures to achieve adequate strength, whereas calcium-rich slag systems react more rapidly under similar conditions [64]. These distinctions in reaction pathways underpin many of the performance outcomes discussed in later sections, making it clear that geopolymerization cannot be understood without reference to the precursor and processing context.

3.2. Mechanism of Geopolymerization

Understanding the key process in the reaction kinetics of geopolymers is essential for setting behavior, bulk microstructure, and property development of geopolymeric gels. Thus, Figure 4 presents a highly simplified reaction kinetic of geopolymerization (according to the main pioneers of the geopolymer system). Though the figure is presented linearly, these reaction mechanisms are largely coupled and occur concurrently, making the basic understanding of the system difficult and complex. Notably, the actual mechanism of setting and hardening of the GPC material is still not clear due to its rapidity. The most proposed and accepted mechanism consists of the chemical reactions that occur in three stages [57,65,66]: (a) dissolution of the aluminosilicate source through the action of hydroxide ion from alkali activators to form reactive precursors (monomers); (b) restructuring, or rearrangement, or condensation of precursor ions to produce oligomers; (c) gelation/polycondensation of oligomers to form polymeric structures.
The high pH of the alkali solution (such as sodium or potassium hydroxide solutions) initiates a dissolution reaction of the aluminosilicate material, leading to the formation of monomers of aluminate and silicate species. This reaction can be apportioned into two periods: (I) dissolution—hydrolysis, and (II) hydrolysis—polycondensation; which are believed to proceed almost simultaneously as the precursor material is mixed with the alkaline solution [67]. Notably, the dissolution of aluminosilicate particles at the surface has always been assumed to be the mechanism responsible for the conversion of solid particles during geopolymerization. As shown in Figure 3, water is usually consumed at this stage, followed by the release of water in other succeeding geopolymerization processes [68,69,70].
After dissolution, the aluminate and silicate species are integrated into the aqueous phase that may react with silicate ions present in the alkaline solution. The complex mixture of aluminate, silicate, and aluminosilicate species is thereby formed. At high pH, the amorphous aluminosilicates dissolve very rapidly, creating a supersaturated aluminosilicate solution. These highly saturated solutions result in the formation of a gel-like structure, and because of condensation, large networks are formed by oligomers/oligo-sialates in the aqueous system [67]. This process also results in the release of water (that was technically consumed during dissolution). As such, water plays an important role in the reaction medium that helps in the formation of a hydrated gel structure. This gel structure is termed bi-phasic, that is, water and aluminosilicate binder form the two phases. The conversion time between the aqueous supersaturated aluminosilicate solution and gel is hypothesized to be dependent on the source material composition, concentration of the alkali solution, and synthetic conditions [71,72]. The processes of rearrangement and reorganization of aluminosilicate bonds determine the microstructure features, pore size, and distribution of the materials, which in turn determine many chemical and physical properties of geopolymeric materials. This stage also leads to the dissolution of more precursor materials, and the solution becomes saturated.
After gelation, and as the linkage of the gel network grows, the system continues to reticulate and reorganize, resulting in the three-dimensional aluminosilicate network commonly regarded as geopolymers—GPC solidification [23,57,66]. The dissolution/alkalination of the aluminosilicate source material and formation of the polymetric species phase are regarded as the nucleation stage, which explains the first two steps proposed by Glukhovsky [73]. This phase is dependent on the thermodynamic and kinetic parameters (such as synthesis of temperature and mixing). While the gel formation, polycondensation, and reorganization phase, in which the nuclei reach a critical size, and crystals begin to mature, is regarded as the growth phase. Although water is not incorporated into the final aluminosilicate network, it plays a critical role throughout the reaction. During the dissolution stage, water facilitates ion transport and enhances the mobility of silicate and aluminate species, while during polycondensation and gelation, water is progressively expelled. The amount and distribution of residual water strongly influence pore structure, strength development, and durability, with excess water often linked to higher porosity and insufficient water associated with incomplete dissolution [74]. This complexity continues to fuel ongoing debate in the literature regarding the precise role of water in geopolymerization.

4. Physio-Chemical Characterization of GPC

4.1. Chemical Properties of FA-Based GPC

While the chemical and microstructural analysis of GPC has been extensively studied, this section uniquely addresses the chemical properties of FA-based GPC in direct comparison to slag-based systems, a gap not previously explored. Various advanced characterization techniques, including X-ray fluorescence (XRF), X-ray diffraction (XRD), Fourier-transform infrared spectroscopy (FTIR), scanning electron microscopy (SEM), transmission electron microscopy (TEM), and thermogravimetric analysis (TGA), have been employed to assess the chemical composition, mineralogy, and microstructural behavior, as well as to predict geopolymerization kinetics and reaction products. This section provides a comprehensive review of these methodologies, highlighting their role in advancing the understanding of GPC chemistry. For instance, XRF is utilized to determine the elemental composition of raw materials, while XRD identifies the crystalline phases present in the GPC reaction products. FTIR provides molecular-level insights by distinguishing different bond types within the GPC network, facilitating both qualitative and quantitative structural analysis. SEM, coupled with energy-dispersive spectroscopy (EDS), enables the examination of micro-to-nanoscale morphology, elemental distribution, and structural characteristics of the synthesized GPC matrix. TEM offers high-resolution phase and elemental analysis, diffraction studies, and real-time observation of dynamic behaviors at the atomic scale. TGA is employed to monitor mass decomposition as a function of temperature and time, providing critical insights into the pozzolanic reactivity and thermal stability of GPC reaction products.

4.1.1. X-Ray Fluorescence (XRF)

XRF determines the elemental composition of GPC reactants, providing quantitative data on the oxide contents of aluminosilicate source materials. Recent studies have extensively utilized XRF to analyze the chemical compositions of various precursors, offering insights into their suitability for dissolution and geopolymerization processes [75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99]. Figure 5 illustrates the average percentage of oxide composition of various FA-based GPC. The main pozzolanic oxides present are silicon dioxide (SiO2) and aluminum oxide (Al2O3), and the minor oxides are iron oxide (Fe2O3), magnesium oxide (MgO), calcium oxide (CaO), potassium oxide (k2O), sodium oxide (Na2O), titanium oxide (TiO2), and sulfur oxide (SO3), etc. Notably, the combined pozzolanic oxides (SiO2, Al2O3, and Fe2O3) in all reviewed studies exceed 50%, classifying them as pozzolanic materials per ASTM C618 and playing a crucial role in geopolymerization reactions. However, the reaction products in a GPC system vary based on the precursor type. FA- and MK-based GPCs, rich in SiO2 and Al2O3, primarily form N-A-S-H gels, contributing to high strength, whereas slag-based GPCs, with higher CaO content, predominantly generate calcium–silicate–hydrate (C-S-H) phases, influencing mechanical and durability properties [100].
In reference to Figure 5, the mean compositions of SiO2, Al2O3, and Fe2O3 are 52.86%, 26.89%, and 6.91%, with standard deviations of 5.63%, 5.19%, and 2.69%, respectively. The minor oxides, including MgO, CaO, K2O, Na2O, and SO3, have mean compositions of 1.85%, 4.27%, 1.43%, 0.47%, and 0.63%, with standard deviations of 1.61%, 3.24%, 0.93%, 0.35%, and 0.50%, respectively, generally ranging between 0.30% and 4%. According to the reviewed studies, the loss on ignition (LOI) varies between 0.04% and 3.89%, primarily due to inefficient and uncontrolled calcination of FA production, which significantly influences geopolymerization and strength development [101]. If the FA has high LOI (unburned carbon content), it will absorb the activator solution, resulting in poor workability of the concrete, a decrease in compressive strength, and ultimately higher cost of production [102].

4.1.2. X-Ray Diffraction (XRD)

XRD is another crucial technique for analyzing GPCs, offering insights into phase composition, microstructural evolution, and material reactivity during geopolymerization. This method is fundamental in distinguishing between amorphous and crystalline phases, enabling a deeper understanding of the transformation mechanisms occurring throughout the synthesis process [75,76,85,103,104,105]. The intensities and diffraction patterns are usually evaluated within the diffraction angle (2Ɵ) between 10° and 80°, allowing for the identification and quantification of both amorphous and crystalline phases and providing insights into the degree of reaction and the nature of the GPC matrix [76,85,103,105].
For a typical FA-based GPC system, XRD analysis usually shows a broad hump in 2Ɵ values between 15° and 35°. The broad hump represents the highly reactive amorphous phase, which is the glassy and disordered aluminosilicate material in the FA that react with alkaline activators during geopolymerization to form the N-A-S-H geopolymeric gel [75,82,85,106]. It also revealed that the major identified crystalline phases are quartz, mullite, and sometimes hematite, which are represented by sharp peaks superimposed on the amorphous hump [60,104,107]. These components originate from the raw FA and do not fully dissolve during geopolymerization, persisting in the final hardened GPC matrix.
To compare the composition of GPC derived from different precursor sources, Figure 6 presents a representative XRD spectrum of FA-based, FA-slag-based, and slag-based GPC matrices, highlighting their distinct mineral compositions [104]. The results confirm that the crystalline phases in FA-based GPC are predominantly represented by sharp peaks of quartz and mullite (alumina silicate), indicative of its high aluminosilicate content [103,107]. In contrast, the slag-based GPC exhibits an amorphous hump at around 28–31° and 10–20°, dominated by high vitreous calcite content, reflecting its high calcium content [108,109]. Consequently, C-A-S-H gels are usually present within the GPC matrix containing slag content [75,110]. The blended FA–slag GPC generally reflects the combined characteristics of both precursors, showing intermediate amorphous features associated with simultaneous N-A-S-H and C-A-S-H gel formation [104]. Quantitative phase analysis was conducted through Rietveld refinement using AUTOQUAN version 2.70, and it revealed that the amorphous content increased from approximately 80% in 100% FA mixes to about 85–86% in slag-rich GPCs, while crystalline phases such as quartz and mullite decreased correspondingly. The higher amorphous fraction in the slag-based matrix indicates greater reactivity and earlier strength gain from rapid C-A-S-H gel formation, thus exhibiting denser microstructures and better mechanical properties [104]. As such, XRD analysis is crucial in assessing the extent of geopolymerization, determining precursor material reactivity, and characterizing the composition of the final GPC product.

4.1.3. Fourier-Transform Infrared Spectroscopy (FTIR)

Among the diverse characterization techniques employed to analyze bond structures in a geopolymer system at the molecular level, only a select few are unique and have been found effective. FTIR is one of the most prominent methods, offering both qualitative and quantitative insights. It has been extensively utilized to investigate the structural evolution and mineralogical composition of FA-based GPC systems [75,80,82,85,111,112,113]. The technique provides insight into the molecular structure and bonding by correlating specific wavenumbers, which correspond to distinct molecular bonds and structures within the materials, to vibrational modes associated with them, and the wavenumber ranges from 4000 to 400 cm−1 [110,114,115].
For a typical FA-based GPC system, FTIR spectra show the formation of bands between 1000 and 1200 cm−1, which are attributed to the asymmetric stretching vibrations of Si-O-T (where T could be Si or Al), and this confirms the development of aluminosilicate frameworks [110,111,112,114,116]. It also reveals the carbonate bands around 1400–1600 cm−1 which are linked to the CO32- group stretching vibrations, indicating carbonation due to the presence of unreacted particles and porous nature [110,114]. FTIR analysis also shows the location of broad bands around 3400–3500 and 1600–1700 cm−1, corresponding to O–H stretching and H–O–H bending of adsorbed H2O molecules, respectively [114,116,117,118]. Other prominent bands revealed by FTIR include the symmetric stretching and bending of Si-O-Si/Al-O-Si (600–800 cm−1) and Si-O bending (400–600 cm−1) [119,120,121].
To compare the composition of GPC derived from different precursor sources, Figure 7 shows the FTIR analysis of FA-based, slag-FA-based, and slag-based GPC matrices, showing various bands with different broadness and intensities [122]. The Si-O-T band (around 1000 cm−1) in FA-based GPC is broader and less intense compared to the slag-based system. This is due to the partial dissolution of FA particles, the amorphous structure, and the slower geopolymerization process [82,112,123]. The carbonate band around 1400 cm−1 is more pronounced in the slag-based system due to the higher calcium and magnesium content, and is present in others owing to the porosity of FA particles [124,125,126]. The O–H stretching (~3400 cm−1) and H–O–H bending (~1600 cm−1) bands are present across all GPC samples, indicating the presence of water. However, these bands appear broader in the FA-based system, suggesting higher porosity compared to the slag-based system, which exhibits a more compact and denser microstructure [117,127,128]. Therefore, FTIR is crucial in the study of chemical properties of different GPC systems, as it provides valuable insight into the microstructure, bonding, and degree of polymerization.

4.1.4. Scanning Electron Microscope (SEM)

SEM analysis provides critical insights into the microstructure of cementitious materials, including GPC, by revealing morphology, particle size, and surface texture, which are key factors influencing material performance. FA-based GPC particles are widely reported to exhibit a predominantly spherical shape with smooth surfaces, which enhances their reactivity and facilitates efficient geopolymerization upon activation with an alkaline solution [106,114,129,130,131,132,133,134,135,136,137]. For GPC mixes made with FA, the spherical nature of the particles aids in creating a firm microstructure, which lowers voids and improves the material properties, and the smoothness of the particles encourages the creation of geopolymeric gels, thus improving the durability and strength of the GPC matrix [138,139]. Many studies also revealed that SEM analysis shows the dispersion of unreacted FA particles in the GPC matrix [60,137,140]. The unreacted particles and cracks are often seen in poorly cured samples, and this can lead to a reduction in the long-term durability performance of the GPC [140]. However, well-reacted FA particles have a more compact microstructure with less porosity, which improves their mechanical performance [141]. Several studies combine SEM with Energy Dispersive X-ray Spectroscopy (EDS) to verify the existence of the aluminosilicate gels and track their formation by detecting changes in chemical composition over time [75,114,130,135,137,142,143]. The microstructural changes identified by SEM-EDS, such as the development of these gels and the quantity of the unreacted particles, are strongly linked to their durability and mechanical properties [144,145]. Essential elements such as Si, Al, Na, and trace amounts of Ca are also confirmed by EDS analysis, which reflects the composition of FA and the products of geopolymerization [142,146].
Typical SEM micrographs of the FA-based (100% FA) and blended FA-slag-based (60% FA–40% slag) GPC matrices from [147] are shown in Figure 8. In the FA-based GPC (Figure 8a), a significant number of unreacted and semi-reacted FA particles are visible, along with the formation of N-A-S-H gel surrounding the particles. This results in a relatively porous microstructure that can reduce early-age strength and durability performance. In contrast, the blended FA-slag sample (Figure 8b) shows the presence of C-A-S-H gel, which forms a denser and more cohesive matrix [148,149]. The angular morphology of the slag particles promotes faster geopolymerization, improving matrix compaction and reducing porosity. However, the higher calcium content and faster reaction rate can also lead to microcracking and shrinkage, which may compromise long-term dimensional stability [148,150,151]. These microstructural differences explain the generally higher mechanical and durability properties of blended FA-Slag GPC compared to 100% FA-based mixes.

4.1.5. Transmission Electron Microscope (TEM)

TEM experiments are a powerful tool that allow for imaging material nano- and microstructure and characterizing material chemistry, revealing details on the distribution and interaction between the crystalline and amorphous phases at high resolution [75,85,110,138,152,153,154]. It operates by capturing and producing high-resolution images that reveal the internal microstructure of the material, after a highly focused electron beam is directed through the thin GPC sample [107,155]. Recent advances in data collection technology for TEM have enabled high-volume and high-resolution data collection at a microsecond frame rate. Thus, this makes TEM the perfect technique for real-time acquisition and microstructural characterization during GPC synthesis (during active dissolution, particle-to-gel conversion, and the formation of hydration products). Previous research that incorporates TEM hypothesized that the key processes occurring in the transformation of the solid aluminosilicate source into the synthetic alkali aluminosilicate are somehow mechanistic and proceed linearly [71,72].
Studies revealed that TEM analysis of FA-based GPC showed the significant presence of nanocrystalline phases—mullite and quart—and amorphous aluminosilicate gels, which are essential in the development of strength and durability of the GPC [59,75,110]. It also shows the transformation of spherical FA particles to homogeneous N-A-S-H gel during geopolymerization by revealing how these particles diffuse to form a compact and dense structure, in which a few weakly crystalline particles due to unreacted FA particles are created together with the GPC gel product [75,110,152]. For example, Figure 9 shows TEM of the FA-based GPC matrix, which shows the amorphous GPC gel formed after the activation of the aluminosilicate and the formation of crystalline materials, which could be quartz or mullite [110,152]. TEM reveals nanoscale details that are important for further understanding and for optimizing these materials for a variety of applications, making it an indispensable tool.

4.1.6. Thermogravimetric Analysis (TGA)

TGA is a useful technique for examining the thermal stability and decomposition behavior of the GPC matrix as a function of weight loss with changes in temperature [115,116,118,156,157,158,159,160,161,162,163]. The rate of weight loss with the corresponding microstructural composition in the FA-based GPC system is mainly studied using TGA within the range of 0–1000°, but most of the weight loss occurs by 700°, and the GPC would be in a stable thermal state [118,164,165,166,167].
Studies revealed that weight loss in FA-based geopolymer typically occurs in three stages [116,160,161,162,168]. The greatest weight loss is usually from the first stage (<200 °C), which is due to the loss of evaporable and residual water within the GPC matrix that is not chemically bound [156,164,167]. The second stage of weight loss occurs between 200 and 400 °C, which is ascribed to the release of chemically bound water from the N-A-S-H gel and the process is known as the dehydration of the hydroxyl group (dihydroxylation) [114,165,169].
The third stage is the carbonate decomposition, which occurs between 400 and 700 °C and usually results in minimal loss in weight [106,166,170]. Öz, Doğan-Sağlamtimur et al., and Aziz, Al Bakri Abdullah et al., presented the TGA thermogram of FA-based and slag-based GPC matrices from [171] and [172], respectively, as shown in Figure 10. It was reported that FA-based GPC mix exhibited lower overall weight loss (about 8%), which could be due to its stable aluminosilicate structure that can better withstand thermal and vapor release during decomposition, compared to the slag-based mix, which experienced about 20% loss in weight at 1000 °C, owing to its richness in calcium compounds that are less thermally stable [112,173,174].
Across the different techniques, consistent patterns emerge between FA- and slag-based systems. XRD and FTIR highlight the dominance of N-A-S-H gels in fly ash mixes and the greater presence of C-A-S-H phases in slag-rich mixes, findings that SEM and TEM confirm through differences in porosity and matrix density. TGA further supports these distinctions by showing lower weight loss and greater thermal stability in FA-based systems compared with the higher mass loss observed in calcium-rich slag mixes. Collectively, these results explain why slag systems often exhibit higher early strength and denser microstructures, while FA systems tend to develop more slowly but offer superior long-term stability.

4.2. Physical Properties of FA-Based GPC

Similar to the chemical analysis of GPC, the evaluation of its physical properties, those measurable without altering the material’s composition, is critical to the performance and durability of GPC, essentially, the net-zero GPC mix. Density, water absorption, porosity, and shrinkage are common tests employed to evaluate the physical properties of GPC [14,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210]. Notably, these properties are influenced by a range of factors, including aggregate characteristics (texture, moisture, shape, and distribution), the ratio of alkaline activators to binder (A/B), sodium hydroxide molarity (SH molarity), the ratio of sodium silicate to sodium hydroxide (SS/SH), curing conditions, the fineness and chemical composition of FA, as well as the incorporation of additives or fibers.
According to research, aggregates have significant impacts on the physical properties of any GPC or mortar, affecting its density, porosity, water absorption, and others. Dense aggregates improve GPC density and enhance thermal insulation [211,212]. Well-graded aggregates minimize porosity while poorly graded or irregularly shaped aggregates increase porosity, which in turn leads to weaker strength [24,212]. Beyond aggregates, which have been extensively studied, this section also predicts the main controlling factor in a GPC system, enabling optimized mix designs. Figure 11 shows the analysis of factors (aggregates excluded) affecting the physical properties of FA-based GPC. The figure demonstrates that the SS/SH ratio is the most influential factor across studies, followed by SH molarity and curing conditions. This underscores the critical role of alkaline activator composition and concentration in determining the physical properties of FA-based GPC systems. SS/SH directly determines the silica content available for gel formation during geopolymerization, and higher SS/SH tends to produce a denser GPC matrix with reduced porosity and absorption [185,191]. Optimizing the SS/SH ratio, which typically ranges between 2.0 and 2.5 depending on precursor chemistry, activator modulus, and related processing conditions, is essential for achieving desirable physical properties [14,213]. Additionally, SH molarity controls the alkalinity of the activating solution and determines the rate of dissolution of the aluminosilicate from FA. Higher SH molarity improves dissolution, thereby enhancing geopolymerization, but may lead to brittleness and shrinkage if excessively high [182,214]. Elevated temperature curing also accelerates geopolymerization, improving density and reducing porosity; however, excessive curing may result in shrinkage [177,190]. Other factors, including the chemical compositions and particle sizes of the aluminosilicate source, also influence density and porosity. Higher surface area and amorphous content increase reactivity, which results in better geopolymerization, reducing porosity and enhancing density [208,209,210].

5. Fresh and Hardened Properties of FA-Based GPC

This section reviews the fresh and hardened properties of FA-based GPC, including workability, compressive strength, tensile strength, and modulus of elasticity. It provides a detailed analysis of the testing procedures and identifies the key factors influencing both the fresh-state behavior and mechanical performance of the system. While numerous review studies and independent research have extensively examined the fresh and hardened properties of GPC mixtures, this section provides a comprehensive analysis of the key factors within the GPC binder that influence these properties—an aspect that remains underexplored in the existing literature. The comparative results presented in this section were derived through a quantitative frequency-based synthesis rather than a formal meta-analysis. Due to inconsistencies in mix designs and reporting formats across studies, statistical pooling was not feasible. Instead, factors influencing each property were identified within individual studies, and their frequency of dominance was counted across all studies and converted into percentages to show how often each parameter was reported as the most influential. This approach provides a clear overview of prevailing trends in the properties despite methodological variations.

5.1. Workability

The workability of FA-based GPC refers to its ease of placement, compaction, and finishing while maintaining cohesion and preventing segregation. Studies have shown that several factors influence the workability of FA-based GPC, including the ratio of alkaline activators to binder (A/B), sodium hydroxide molarity (SH molarity), sodium silicate-to-sodium hydroxide ratio (SS/SH), admixtures such as superplasticizers, and the fineness and chemical composition of FA. Additionally, parameters such as solid-to-liquid ratio, water-to-binder ratio, and the incorporation of slag also play significant roles in determining the rheological behavior of the system [14,25,91,201,213,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235,236,237,238,239,240,241,242,243,244].
FA-based GPC generally has better workability than slag-based GPC due to the spherical shape of fly ash particles, which reduces friction and enhances flowability [130,243]. In contrast, the irregular and angular particles in slag-based GPC increase water demand, leading to rapid stiffening [135,245]. Additionally, FA has a slower reaction rate, allowing for extended workability and easier handling, whereas slag is highly reactive, accelerating setting times and consequently reducing workability over time [239,246].
Figure 12 presents a synthesis of factors influencing workability based on the analysis of 35 studies. SH molarity was the most frequently investigated parameter (n = 17), followed by SS/SH ratio (n = 16) and A/B ratio (n = 15). However, when evaluating relative impact, the A/B ratio emerged as the most influential factor. Among the studies that assessed at least two of these three parameters (SH molarity, SS/SH ratio, and A/B ratio), 71.4% identified A/B as having the greatest effect on workability, while 28.6% favored SH molarity, and none identified SS/SH as the dominant influence. This finding highlights the critical role of A/B ratio in determining workability, as it directly controls the amount of liquid available in the mix. A higher A/B ratio generally results in improved flow due to increased activator volume, but beyond a certain threshold, it may cause segregation or bleeding [14,240]. SH molarity also significantly impacts workability. Lower molarities increase the water content in the activator solution, reducing viscosity and enhancing flow. Higher molarities, in contrast, accelerate dissolution and early geopolymerization, increasing stiffness and reducing workability [213,217,220,223,232]. While the SS/SH ratio affects the viscosity of the activator solution and the mix’s consistency, no study in this review identified it as the most decisive factor for workability. Nevertheless, it remains an important secondary parameter influencing fresh mix behavior [14,201,213,229].

5.2. Hardened Properties

Compressive strength, tensile strength and modulus of elasticity are significant engineering properties, of FA-based GPC systems, which have been widely studied, and numerous studies have shown that factors like A/B ratio, SH molarity, SS/SH ratio, additives (slag content, nano-silica), curing conditions, and many others (aggregate presence and extra water) have an influence on them [14,60,81,88,105,200,213,215,217,223,229,232,236,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267].
FA-based GPC generally exhibits lower compressive strength compared to slag-based GPC due to its lower calcium content, slower reaction rate, and more porous microstructure [243,266]. Slag contains a higher amount of calcium (CaO), which promotes the formation of C-S-H and C-A-S-H gels, significantly enhancing strength [243]. In contrast, FA is primarily aluminosilicate-based and relies on alkali activation to form N-A-S-H gel, which provides strength but is generally weaker than C-S-H [246,268]. Additionally, slag-based geopolymer develops a denser and more compact microstructure due to increased C-S-H gel formation, while FA-based GPC tends to be more porous, which can slightly reduce its compressive strength [112,245].
Among the 35 studies reviewed for the compressive strength of FA-based GPC, many investigated multiple parameters simultaneously. Figure 13 presents a synthesis of the number of studies considering each factor, alongside the proportion of studies identifying a particular factor as the most influential on compressive strength. SH molarity was the most frequently studied parameter (n = 23), followed by SS/SH ratio (n = 20) and A/B (n = 15). However, when evaluating relative impact, the A/B ratio emerged as the dominant factor, with 70.6% of the studies that considered at least two of the three factors (SH molarity, SS/SH ratio, and A/B ratio) identifying it as the most influential. SH molarity was dominant in 23.5% of studies, while SS/SH ratio was most influential in only 5.9% of the studies. An increase in the A/B ratio generally enhances compressive strength; however, exceeding the optimal threshold may lead to a reduction in strength. This reduction is due to an excessive concentration of the alkaline activator, which disrupts the formation of a stable GPC matrix, compromising structural integrity [239,240,263].
Ghafoor et al. [14] reported that increasing the A/B for fly ash-based GPC from 0.4 to 0.5 resulted in a reduction in compressive strength for cases composed of SH molarity 8 M and 10 M. A similar trend was also found by Arafa et al. [217], where around 10% reduction was observed when the A/B ratio was increased to 0.5 for a 10 M SH-based GPC compared to an A/B 0.4 mixture. However, literature indicates that the optimum A/B is dependent on the type of precursor used; more specifically, the particle size distribution of the precursor is the influencing factor [221]. The development of compressive strength of GPC depends on the dissolution of silicon and aluminum particles in the geopolymerization process, and the reaction is dependent on the molarity of the alkaline activator (e.g., SH). To understand the effect of the concentration of the activator, several studies have been conducted, and it was observed that increasing the molarity of the activator (e.g., NaOH) positively influences strength development [14,217]. However, there is also an optimum level of 14 M reported by some researchers, where they speculated that at higher molarity, there might be a congestion of hydroxide ions (OH) in the matrix [14]. Alongside A/B and SH molarity, another influencing factor for GPC is the SS/SH ratio. The SS supplies the Si4+ ions into the mix to accelerate the reaction and help in building long, stable polymeric chains. Similar to the A/B ratio, increasing the SS/SH ratio also adversely affects the compressive strength of GPC beyond a certain point. However, the optimum SS/SH ratio is system-dependent.
The curing conditions, which include temperature and environment (ambient curing, steam curing, oven curing), also have a crucial impact on the development of FA-based GPC strength. The geopolymerization process is accelerated by high curing temperatures, which produce denser and more compact materials, and result in higher early strength of GPC mixes’ compressive strength [258,259,262]. However, the required strength could also be reached with ambient curing, but it takes longer than heat curing [247,254,255].
Compressive strength is considered the most substantial mechanical property of concrete also the most crucial parameter to initiate any construction project, and it is observed that other mechanical properties are primarily dependent on it. Typically, the tensile strength of GPC follows a similar trend to its compressive strength, where an increase in compressive strength is accompanied by an enhancement in tensile strength [269,270,271]. Similar to compressive strength, the tensile property of the GPC system also depends on A/B ratio, SH molarity, SS/SH ratio, curing conditions, and the incorporation of additives [24,60,84,86,91,142,150,200,201,202,227,237,238,240,242,254,272,273,274,275,276,277,278,279,280,281,282,283,284]. However, tensile behavior is also uniquely influenced by the quality of the matrix–aggregate interface and the binder’s ability to transfer stress across microcracks. Fiber incorporation (such as steel, polypropylene, or basalt) enhances this behavior by bridging cracks, delaying their propagation, and improving ductility and energy absorption [272,273,274,276]. The inclusion of additives and reinforcing fibers also contributes to the tensile strength properties by modifying the microstructure, thereby improving the tensile strength, which in turn improves cracking behavior, ductility, and tensile strength of FA-based GPC [86,272,273,281,285].
The modulus of elasticity (MOE) is another important parameter in assessing the deformation behavior of FA-based GPC, and studies have shown that similar factors such as A/B ratio, SH molarity, SS/SH ratio, curing conditions, additives, and the inclusion of fibers also influence its performance [14,43,142,237,238,240,241,242,282,286,287,288,289,290,291,292,293,294,295,296,297,298,299,300,301,302,303]. Additives enhance the MoE of FA-based GPC by increasing gel density, stiffness, and improving microstructural integrity through the formation of denser N–A–S–H networks [287,297,298,299]. In contrast to tensile strength, MoE is strongly affected by aggregate stiffness, gel composition, and porosity. Higher degrees of N–A–S–H polymerization and reduced pore volume yield greater stiffness, whereas elevated porosity and weak interfacial transition zones reduce the modulus. Thus, while compressive strength remains a general indicator, tensile and elastic behaviors are independently controlled by microstructural cohesion, gel chemistry, and fiber–matrix synergy.
In general, optimal workability of the GPC matrix can be achieved without compromising strength by maintaining an A/B ratio between 0.4 and 0.6, SH molarity between 10 and 14 M, and an SS/SH ratio of 1.5 to 2.0. Curing conditions also play a role in the strength development of FA-based GPCs, particularly in precast applications. Early curing at temperatures between 40 °C and 60 °C for 24–48 h, typically followed by ambient curing, maximizes strength and MoE; however, this approach demands significant energy, potentially increasing the carbon footprint. Conversely, ambient temperature curing is a more sustainable and practical alternative for in situ applications, but it results in lower early-age strength and a longer time to achieve the desired strength and MoE. Enhancing early-age strength can be accomplished by incorporating additives that accelerate geopolymerization and extending the curing period, allowing for competitive strength and MoE with minimal energy consumption. In large-scale applications, however, maintaining uniform temperature and moisture during this hybrid curing process remains a practical challenge.

6. Durability Properties of FA-Based Geopolymer

This section reviews the durability-related properties of FA-based GPC, including acid attack, chloride attack, sulfate attack, and chloride-induced corrosion [304,305,306,307,308,309,310,311,312,313,314,315,316,317,318,319,320,321,322,323,324,325,326,327,328,329,330,331,332,333,334,335,336,337,338,339]. This section presents a detailed analysis of the effective testing methodologies currently employed to assess the long-term durability, performance, and serviceability of GPC structures. Additionally, a comparative evaluation of the durability properties of FA-based and slag-based GPCs against OPC concrete is conducted. A summary of strategies for optimizing GPC durability is also provided. Building on the evaluation of fresh and hardened properties, this section offers a comprehensive assessment of the critical factors within the GPC binder that influence durability, including chemical composition, SS/SH ratio, curing conditions, and the incorporation of additives. Despite the growing body of research on GPC, the impact of these parameters on long-term durability remains insufficiently explored, making this review a significant contribution. Figure 14 represents the performance matrix for GPC and OPC concrete against several aggressive agents.

6.1. Resistance to Acid Attack

The GPC system has shown remarkable potential for resistance against acid attack due to its chemical composition and microstructure. The ASTM C1898 test method is preferentially used to assess the acid resistance of concrete and mortar by immersing specimens in acidic solutions and measuring mass loss, strength reduction, and surface deterioration over time, with lower degradation indicating higher durability [340]. The low calcium content in the FA-based GPC system results in lower quantities of gypsum and ettringite formation, which tends to deteriorate under acidic attack [316,317,321]. In contrast to OPC concrete, which exhibits significant weight loss and strength reduction in acidic environments, studies have shown that the GPC matrix maintains its structural integrity under such conditions [307,314,315]. Similarly, slag-based GPC showed resistance to chloride attack but was slightly inferior to FA-based GPC due to the presence of calcium phases that form gypsum and ettringite, which may weaken the structure under acidic conditions. Additionally, the incorporation of additives (including nanosilica and silica fume) improved the resistance to acid attack due to their ability to modify the geopolymer matrix to a denser one, thereby minimizing degradation [74,135,318]. Higher molarity of SH between 10 and 12 M creates a denser geopolymer matrix with better resistance to acid attack [306,313]. Likewise, elevated temperature curing also enhances geopolymerization, resulting in reduced porosity and improved acid resistance [319].

6.2. Resistance to Chloride Attack

Another significant long-term property of GPC in chloride-rich environments is the material’s resistance to chloride attacks. This is typically assessed using the Rapid Chloride Permeability Test (RCPT), which quantifies the electrical charge passing through a 50 mm thick, 100 mm diameter concrete specimen over a 6 h period under 60 V DC voltage. The measured charge, expressed in coulombs, serves as an indicator of the material’s resistance to chloride ion penetration, with lower values signifying higher durability. This test is crucial for evaluating the long-term performance of concrete in chloride-rich environments [341].
FA-based GPC has significantly less permeability compared to OPC concrete due to the dense microstructure of its geopolymer matrix that reduces the ingress of harmful ions into it [307,311,342]. Similarly, slag-based GPC showed excellent resistance to chloride attack due to a denser microstructure and lower permeability, resulting from the formation of C-A-S-H. Moreover, the chemical bonding in the GPC structure adds to its durability regarding chloride-induced degradation, thus making it fit for marine and coastal applications as well [305,326]. Studies have shown that FA-based GPC performs better in a chloride environment, since the degrading effect observed in OPC concrete is not recorded for it [327,343]. SS/SH plays an important role in the formation of denser geopolymer matrices, and the inclusion of additives (such as slag, nano-titanium dioxide) modifies the microstructure, which makes it highly resistant to chloride attack [135,308,325].

6.3. Resistance to Sulfate Attack

The GPC system also demonstrates good resistance to sulfate attack due to its aluminosilicate network, which is less reactive to sulfate ions. ASTM C1012 [344] is usually used to assess the sulfate resistance of cementitious materials by measuring the length expansion of mortar or concrete bars immersed in a sodium sulfate (Na2SO4) solution over an extended period (e.g., 6 months to 1 year). Lower expansion indicates higher sulfate resistance, making this test critical for evaluating the durability of concrete in sulfate-rich environments.
The polymeric N-A-S-H gel structure in FA-based GPC minimizes the formation of expansive products like ettringite, reducing sulfate-induced deterioration and enhancing durability in sulfate-rich environments [317,320,328]. Indeed, in the literature, it is reported that FA-based GPC exposed to sulfate solutions showed only a slight reduction in strength after an extended exposure period, indicating its good performance in such conditions [321,330,345]. For slag-based GPC, studies show moderate resistance because sulfate ions react with calcium phases in slag, leading to the formation of expansive ettringite and gypsum. Additionally, C-A-S-H phases can interact with sulfates, forming expansive products, which may compromise durability. Allahverdi et al. [346] reported that slag-based GPC showed 0.03% expansion, whereas Saavedra et al. [323] reported an expansion of 0.0048% for FA/slag GPC, where FA was the dominant precursor. Notably, the addition of alternative materials, such as nanosilica, silica fume, and rice husk ash, significantly improve their sulfate attack resistant by reducing porosity and densifying the GPC matrix, thereby limiting the formation of expansive sulfate reaction products [322,323,324].

6.4. Resistance to Corrosion

Corrosion resistance is a critical concern for reinforced concrete structures, especially when exposed to harsh environmental conditions such as moisture, oxygen, carbon dioxide, chlorides, and sulfates. The Half-Cell Potential Test is among the most widely used non-destructive techniques for evaluating the likelihood of corrosion in reinforced concrete. This method measures the electrical potential of embedded steel using a copper–copper sulfate electrode, where more negative readings indicate an increased risk of corrosion, aiding in the early detection and mitigation of deterioration [347]. The absence of Portland cement in FA-based GPC and the formation of an alkali-activated aluminosilicate matrix influence the corrosion resistance properties, protecting embedded steel reinforcement [331,332]. In chloride-rich environments, GPC has demonstrated lower corrosion rates and prolonged durability than OPC concrete structures. A study by Reddy et al. (2013) assessed the resistance of GPC to chloride attacks and found that it exhibited excellent resistance to corrosion-induced cracking, with longer times to corrosion initiation compared to OPC concrete [333]. Similarly, different studies also found that FA-based GPC exhibited low chloride permeability, indicating reduced corrosion potential [335,338]. Slag-based GPC exhibits superior resistance due to its denser microstructure, which reduces the ingress of corrosive ions. The composition and type of activators used in GPC play a crucial role in its corrosion resistance [348,349]. Studies have indicated that SS/SH improves the binding properties of the GPC system, leading to better corrosion resistance [337]. Moreover, the incorporation of slag along with FA enhances the GPC matrix, reducing its porosity and improving durability [334].
Figure 14 and Table 1 offers a comparative summary of the durability properties of OPC concrete, FA-based GPC, and slag-based GPC, highlighting the fundamental science behind these materials and tailored strategies for enhancing resilience in aggressive environments. Remarkably, GPC mixtures outperform OPC in resisting acid attack, chloride penetration, sulfate exposure, and corrosion, making them a more durable choice for harsh conditions. However, optimizing critical mix design parameters remains essential to unlocking their full potential and ensuring long-term structural integrity.
To provide a deeper understanding of the factors influencing GPC durability and highlight its potential for enhanced performance in aggressive environments, interrelationships among various durability performance properties were analyzed. Figure 15a,b illustrates the correlation between resistance to acid, sulfate, and chloride attacks in GPC, as observed and synthesized from experimental datasets in the literature. The trendline equations (y = 0.7883x + 23.745 and y = 0.2271x + 76.526) with R2 values of 0.8833 and 0.774, respectively, demonstrate a good relationship and a strong positive correlation among durability properties. This suggests that when GPC shows increased resistance to one type of chemical attack, such as acid, its resistance to sulfate and chloride attacks also improves, though at varying levels. This interdependence arises from the chemical and microstructural evolution of geopolymer binders. A higher degree of geopolymerization produces a denser and more continuous reaction gel, typically N–A–S–H in low-calcium systems and C–A–S–H or hybrid N–(C)–A–S–H in calcium-rich systems, that fills pores and refines the pore structure [317,320,328]. The resulting compact matrix reduces permeability and ion diffusivity, thereby limiting the ingress of aggressive species and suppressing secondary reactions such as gypsum or ettringite formation. Consequently, a well-reacted, low-porosity matrix exhibits improved resistance to multiple deterioration mechanisms simultaneously. Ultimately, GPC exhibits a strong interdependence between acid, sulfate, and chloride resistance, making it highly suitable for infrastructure in aggressive environments.
However, it should be recognized that most durability investigations on geopolymer concretes have been conducted under accelerated laboratory conditions designed to simulate long-term degradation within shorter periods. While such methods are valuable for comparative evaluation, they often exaggerate the severity of exposure and may not fully represent in-service conditions. Therefore, translating laboratory durability results into reliable field predictions requires additional long-term studies and environment-specific validation.

7. Methodological Quality and Limitations of Reviewed Articles

The assessment of methodological quality of the reviewed articles is important to understand their reproducibility and broader applicability. For most of the investigations, which researched both the mechanical and durability parameters of geopolymer concrete, the studies were small in scale. Moreover, considerable variability was observed among reported results, particularly for mechanical properties such as compressive strength, due to differences in precursor chemistry, activator composition, curing temperature, and testing age. In addition, the results were generally presented as mean values without corresponding measures of variability, such as standard deviation or error ranges, which limits the statistical reliability and comparability of findings.
It has been reported that the strength and quality of GPC are dependent on precursors. Despite this, the majority of the studies relied on single-source fly ash, with chemical composition reported but limited mineralogical and microstructural verifications [14,213,217,237,238,241,250,321,325].
However, there are studies that compared the effects of precursor variability on both hardened and durability properties [91,93,201,222,227,229,239,240,251,252,256,263,291,311,314,328,348,353], but the absence of thorough characterization made the comparative conclusions unreliable. Alongside the lack of precursor variability, the scope of most of the articles was limited. For instance, most studies focused on early-age strength development, and very few examined the effect of long-term curing (56, 90, and 180 days) on mechanical properties [86,311,328,353]. Although the scope of the studies was limited to short-term analysis for mechanical properties, durability properties such as acid and sulfate resistance, chloride resistance were examined in depth for variables like precursor type, SH concentration, and SH/SS ratios.
Considering the methodological differences across studies, this review identifies prevailing trends and evidence-based patterns that provide valuable insights while recognizing system-specific variations. These findings emphasize the importance of improving experimental consistency across future studies, particularly in reporting mix proportions, activator compositions, and precursor characteristics. Enhancing methodological transparency and data comparability will enable more robust evaluations of mechanical and durability performance across different GPC systems.

8. Cost and Sustainability Analysis of GPC

This section examines the economic and environmental performance of GPC along with its variants, in comparison to the OPC concrete matrix. Life Cycle Assessment (LCA) is commonly employed to evaluate the environmental impacts of both OPC concrete and GPC variants, while life cycle cost analysis (LCCA) is used to assess the economic feasibility of GPC mixtures. Although numerous academic studies have explored the environmental impact of OPC concrete, the LCA and LCCA of GPC variants relative to OPC mixes remain underexplored. Although a comprehensive regional market analysis was beyond the scope of this review, it is important to note that precursor availability and logistics vary significantly across regions. For instance, fly ash remains more abundant in Asia due to continued coal combustion, whereas its supply is declining in the United States and parts of Europe following coal plant retirements. Slag availability is more stable in industrialized regions with active steel production. These regional variations influence both cost and sustainability metrics, highlighting the need for future LCAs and techno-economic studies to consider localized supply chain effects. From the case study analyses, the findings in this section emphasize the importance of selecting appropriate precursor materials and optimizing processing conditions, while also considering regional availability and logistics, to enhance the adoption of GPC as a viable alternative to OPC concrete.

8.1. Regional Variation in Production Cost of GPC and OPC Concrete

To understand the regional variation in the production cost of OPC concrete and GPC mixtures, a simplified cost analysis was performed. For consistency, 1 m3 of concrete with equivalent compressive strength ranging between 40 and 42 MPa was considered as the functional unit. Mixtures with equivalent compressive strength ranges are assumed to have similar functions and, thus, were compared together. However, the main purpose of this analysis was to showcase the regional variability of material costs and thus the variability in the production cost of GPC and OPC. The unit prices of the constituents of GPC and OPC for different regions were collected from the literature [355]. Subsequently, these values were multiplied by the amount of each constituent to obtain the total cost of each mixture and to estimate the regional variations. For this simplified cost analysis, only the prices of distinct constituent materials (i.e., OPC, FA, slag, SH, and SS) were considered as the main variables, while energy cost, labor cost, and transportation cost were not included. Figure 16 shows the regional price variation in OPC and GPC mixtures, and it is evident that, irrespective of the regional variation, GPC mixtures are almost one-and-a-half times as costly as their OPC-based counterparts on average for the markets of the USA and China. From the figure, it is also evident that, for FA-based GPC mixtures, the usage of sodium silicate is the main reason for making them the most costly mixtures. This cost disparity suggests that, while GPC presents significant environmental benefits, its higher production cost could be a major barrier to widespread adoption.
In addition to the above-mentioned simplified cost analysis based on material price, the overall cost associated with the production of 1 m3 of GPC and OPC concrete can be sensitive to the costs associated with transportation of raw materials and production of concrete, which can vary significantly by regional location. It has been reported in the literature that considering the long-distance import of GGBFS surpasses the environmental impacts related to OPC in 15 categories [356]. Therefore, it can be understood that, alongside environmental impacts, the cost of precursors can be influenced by location and transportation requirements.

8.2. Case Studies Sustainability Analysis

Table 2 shows the data statistics, including consumption rate, price and global supply and demand of FA, slag, FA-based GPC, and slag-based GPC. The study shows that FA-based GPC presents the most promising alternative to OPC concrete, given its widespread availability and potential to eliminate a material that poses considerable environmental and health risks. Although the production rate of slag is lower than that of FA, with slag-based GPC replacing just 6.5% of annual OPC production. The analysis, however, indicates that slag is a more favorable option due to its cost-effectiveness, improved workability (minimizing segregation), and improvements in long-term performance. Notably, the high production cost of the activators (SS and SH) remains a challenge for these GPC mixes, requiring further technological advancements to reduce both activator costs and production energy.
The comparative environmental evaluation of OPC and GPC produced from FA and slag, based on recent studies [358,359,360], was investigated. Table 3 presents the sustainability analysis and environmental impact, comparing FA-GPC, slag-GPC, and OPC. The functional unit is defined as 1 m3 of concrete with specified compressive strength. All analyses conducted in each study used cradle-to-gate LCA methodology, considering freshwater and toxicity, climate change, ozone depletion, acidification, energy consumption, and GHG emissions for all phases. As shown in Table 3, the analysis indicates that slag-based GPC has lower environmental impacts than FA-based GPC. A negative percent change signifies a reduction in economic, social, or environmental burdens. For instance, the energy consumption and CO2 emissions for slag-based GPC are 33% and 31% lower than those of FA-based GPC and 62% and 77% lower than OPC mixes. This is because the use of slag reduces reliance on sodium silicate, further decreasing the energy demand for GPC production. In GPC mixes, the higher global warming potential and CO2 emissions compared to OPC are primarily due to the high energy demand of sodium silicate production (the commonplace activator). This solution has major environmental impacts (such as terrestrial ecotoxicity, ozone layer depletion, and human, freshwater, and marine toxicity) on their long-term sustainability benefit [361]. This activating solution also limits its sustainability benefits due to the required high curing heat, difficulties in mixing and field handling, viscous nature, and high cost [362,363]. In a previous study, it was reported that the burden of sodium silicate ranges between 55 and 74% in all impact categories considered in an LCA study of geopolymer blocks, and that the indirect use of SS and the direct use of SH are the main contributors to environmental burdens [364]. In the same study, it was discussed that the intensity of the environmental burden depends on the production process of the activators. The production of NaOH is considered the most energy- and cost-intensive process, and this arguably controls the LCA of the geopolymer concrete system. The other constituent materials affect the environment in their individual ways, while NaOH’s production affects most of them [364]. Arguably, incorporating slag cement can mitigate this impact by reducing the required amount of sodium silicate, thereby lowering overall energy consumption and emissions. Another approach to overcome the impacts related to activator utilization is the development and use of waste-based solid activators in one-part (“just add water”) GPC, which represents a promising alternative, as it can reduce embodied carbon, simplify handling, and enhance practical applicability for large-scale construction.

9. Conclusions and Future Research

9.1. Conclusions

While numerous review papers have been conducted on GPC, there remains a lack of systematic studies comparing GPC made from different precursors, particularly FA and slag. Additionally, recent studies over the past decade have largely overlooked a comprehensive approach that not only examines the engineering properties of GPC but also considers its sustainability aspects. To address these gaps, this review critically evaluates all key branches of the GPC system, including physico-chemical properties, mechanical performance, durability, and sustainability considerations. This study aims to provide a foundation for developing or proposing a net-zero GPC, ensuring both structural efficiency and environmental viability. The key findings from this study are summarized as follows:
  • GPC is synthesized by activating aluminosilicate precursors to produce N-A-S-H and C-A-S-H gels, which govern strength and durability. Low-Ca precursors such as FA predominantly form N-A-S-H, while high-Ca precursors like slag promote C-A-S-H gel formation, resulting in a denser microstructure and higher strength. However, this can also increase the risk of cracking due to the rapid geopolymerization process.
  • Achieving optimal physical properties for net-zero GPC requires careful mix design, with particular attention to the SS/SH ratio. Characterization studies confirm that FA-based GPC primarily forms N-A-S-H gels, while slag-based GPC produces C-A-S-H phases, impacting strength and durability. These differences influence the reaction kinetics, gel structure, and thermal stability of the resulting matrix.
  • The workability and mechanical properties of GPC, including compressive strength, tensile strength, and MOE, are significantly influenced by the A/B ratio. While high-Ca precursors improve mechanical strength due to the denser nature of C-A-S-H gels, they also accelerate setting time, reducing workability. Blending high- and low-Ca precursors and the use of admixtures can enhance workability without compromising strength, making GPC suitable for both precast and in situ applications.
  • GPC exhibits superior resistance to acids, chlorides, sulfates, and corrosion compared to OPC concrete, owing to its aluminosilicate-based matrix that limits the formation of expansive degradation products. The durability of GPC is closely tied to its microstructural composition, and blending high- and low-Ca precursors can enhance its performance in aggressive environments. This strategic approach not only improves long-term durability but also supports the goal of achieving net-zero carbon GPC.
  • The analysis, while assessing global feasibility, identifies FA-based GPC as the most viable alternative to OPC concrete due to its abundant global supply and potential to mitigate environmental and human health risks. However, sustainability assessments indicate that slag-based GPC exhibits lower overall environmental impacts compared to FA-based GPC. Using slag reduces reliance on SS, further decreasing the energy demand for GPC production.
  • The review indicates that, based on precursor availability, curing techniques, and exposure environment, mass-level production of geopolymer concrete is feasible. However, adoption should be gradual—aligned with project-specific needs and progressing toward performance-based applications. In the meantime, alternative precursor materials are needed to address shortages of conventional precursors, and parallel research on new activators should continue.

9.2. Future Research

As GPC continues to emerge as a viable net-zero carbon alternative to OPC, several challenges must be addressed to enhance its sustainability, performance, and large-scale adoption. Future research should focus on optimizing material formulations, improving mechanical and durability properties, and refining sustainability strategies to further reduce the carbon footprint of GPC. The following research directions are essential in achieving these goals:
  • With the declining availability of conventional precursors like FA and slag, future research should focus on identifying and optimizing alternative SCMs to ensure long-term sustainability. Potential sources such as agricultural residues, waste glass, and industrial byproducts offer promising low-carbon alternatives due to their abundance and reactivity. Investigating their geopolymerization kinetics, compatibility with different activators, and long-term performance is critical for developing sustainable, widely available, and cost-effective GPC binders that support carbon neutrality in construction.
  • Optimizing GPC mix design through precise control of A/B and SS/SH ratios for given precursors is crucial to achieving net-zero carbon goals while ensuring high-performance standards. Future studies should focus on a systematic approach that tailors A/B and SS/SH ratios based on the specific precursors used to enhance workability, strength, and durability.
  • While GPC is a more sustainable alternative to OPC, its feasibility for large-scale OPC replacement remains limited in many regions worldwide. Thus, future research should focus on minimizing the use of SH and SS in GPC activators due to their high cost and limited sustainability. The development and use of waste-based solid activators in one-part (“just add water”) GPC represents a promising alternative, as it can reduce embodied carbon, simplify handling, and enhance practical applicability for large-scale construction.
  • Future research should explore emerging innovations such as CO2 curing and AI-driven mix design. CO2 curing has shown potential to enhance early-age strength, densify the GPC matrix, and reduce overall carbon emissions by promoting controlled carbonation. Similarly, artificial intelligence and machine learning techniques can be leveraged to optimize mix design by predicting the performance of GPC based on input variables such as precursor type, activator ratio, and curing conditions. These tools can minimize experimental trial-and-error and accelerate the development of high-performance, net-zero GPC systems.

Author Contributions

Funding and Conceptualization, E.O.F.; methodology, E.O.F. and T.O.M.; software, A.U.H. and M.Z.B.H.; validation, A.U.H. and M.Z.B.H.; formal analysis, T.O.M.; investigation, T.O.M. and E.O.F.; resources, T.O.M. and E.O.F.; data curation, T.O.M., A.U.H., and M.Z.B.H.; writing—original draft preparation, T.O.M.; writing—review and editing, E.O.F. and T.O.M.; visualization, E.O.F. and T.O.M.; supervision, E.O.F.; project administration, E.O.F. All authors have read and agreed to the published version of the manuscript.

Funding

The authors declare that financial support was received for the research and/or publication of this article from the GT Building Team Grant (DE00024085) and the Georgia Tech Sustainability NEXT Grant (DE00023413). This work was performed in part at the Georgia Tech Institute for Matter and Systems, a member of the National Nanotechnology Coordinated Infrastructure (NNCI), which is supported by the National Science Foundation (ECCS-2025462).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available upon request.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

OPCOrdinary Portland cement
GPCGeopolymer concrete
SCMSupplementary cementitious material
FAFly ash
BMTBillion metric tons
CACCalcium aluminate cements
CSACalcium sulfoaluminate cements
MKMetakaolin
BWWABrisbane West Well Camp Airport
XRFX-ray fluorescence
XRDX-ray diffraction
FTIRFourier-transform infrared spectroscopy
SEMScanning electron microscopy
TEMTransmission electron microscopy
TGAThermogravimetric analysis
EDSEnergy-dispersive spectroscopy
LOILoss on Ignition
A/BActivators to binder ratio
SHSodium hydroxide
SSSodium silicate
N-A-S-Hsodium-aluminate-silicate-hydrate
C-A-S-Hcalcium-aluminate-silicate-hydrate
C-S-HCalcium-silicate-hydrate
MoEmodulus of elasticity
RCPTRapid Chloride Permeability Test

Appendix A

Chemical composition of FA from various studies.
ReferenceSiO2Al2O3Fe2O3MgOCaOK2ONa2OSO3TiO2P2O5LOIPO
[75]55.6029.805.911.081.591.940.230.451.63-0.4791.31
[76]52.0620.545.503.2914.070.690.570.57--0.1078.10
[77]53.7028.106.991.594.321.890.87----88.79
[78]57.9031.115.070.971.291.000.090.05--0.0494.08
[79]48.4039.6012.101.302.700.300.200.30- 1.70100.10
[80]37.7224.158.413.652.734.57-1.371.251.01-70.28
[81]61.8527.365.181.001.470.630.080.051.840.541.0094.39
[82]52.7920.957.763.426.950.510.09-0.85--81.50
[83]52.5022.825.342.567.160.990.480.20--3.3580.66
[84]43.7320.1812.373.7511.141.960.931.45---76.28
[85]52.5030.202.941.230.822.08--1.03-7.1285.64
[86]50.7028.808.801.392.382.400.840.30--3.7988.30
[87]53.7033.203.600.503.000.80-0.601.600.402.6090.50
[88]48.9019.6311.564.316.062.060.731.65--2.3280.09
[89]52.1123.597.390.782.610.800.420.490.881.31-83.09
[90]63.3226.765.550.292.490.00020.00040.36--0.9795.63
[91]54.0019.606.906.907.902.20--0.880.341.8780.5
[99]53.7032.905.500.921.841.760.370.462.100.15-92.1
[92]52.8321.5010.490.896.441.760.82-1.601.751.5084.82
[93]58.2325.084.561.212.870.870.411.160.830.201.5987.87
[94]60.4828.154.520.471.711.410.14---1.5993.15
[95]55.9028.106.97 3.841.55--2.21-1.2090.97
[96]44.8329.234.661.624.470.681.320.62---78.72
[97]49.1034.804.500.404.901.300.40---2.3088.4
[98]55.0026.0010.170.802.091.650.40---3.8991.17
PO—Pozzolanic oxide.

References

  1. Gagg, C.R. Cement and concrete as an engineering material: An historic appraisal and case study analysis. Eng. Fail. Anal. 2014, 40, 114–140. [Google Scholar] [CrossRef]
  2. Sandanayake, M.; Bouras, Y.; Haigh, R.; Vrcelj, Z. Current sustainable trends of using waste materials in concrete—A decade review. Sustainability 2020, 12, 9622. [Google Scholar] [CrossRef]
  3. Chen, X.B.; Liang, J.F.; Li, W. Compression stress-strain curve of lithium slag recycled fine aggregate concrete. PLoS ONE 2024, 19, e0302176. [Google Scholar] [CrossRef]
  4. Mohammed, T.O.; Harun, M.Z.B.; Liu, J.; Fanijo, E.O. Effect of Oil Palm Broom Fiber on the Mechanical Properties of Rice Husk Ash–Blended Concrete. Recent Prog. Mater. 2025, 7, 13. [Google Scholar] [CrossRef]
  5. Miller, S.A.; Horvath, A.; Monteiro, P.J.M. Impacts of booming concrete production on water resources worldwide. Nat. Sustain. 2018, 1, 69–76. [Google Scholar] [CrossRef]
  6. Fanijo, E.; Babafemi, A.J.; Arowojolu, O. Performance of laterized concrete made with palm kernel shell as replacement for coarse aggregate. Constr. Build. Mater. 2020, 250, 118829. [Google Scholar] [CrossRef]
  7. Maywald, C.; Riesser, F. Sustainability—The Art of Modern Architecture. Procedia Eng. 2016, 155, 238–248. [Google Scholar] [CrossRef]
  8. Farooq, F.; Jin, X.; Javed, M.F.; Akbar, A.; Shah, M.I.; Aslam, F.; Alyousef, R. Geopolymer concrete as sustainable material: A state of the art review. Constr. Build. Mater. 2021, 306, 124762. [Google Scholar] [CrossRef]
  9. EPA. 2019 Direct Carbon Intensities Quartile Metric Ton CO2/Metric Ton of Clinker Metric Ton CO2/Metric Ton of Cement Carbon Intensity. 2021. Available online: https://www.ecfr.gov/cgi-bin/text- (accessed on 14 September 2024).
  10. Ige, O.E.; Von Kallon, D.V.; Desai, D. Carbon emissions mitigation methods for cement industry using a systems dynamics model. Clean Technol. Environ. Policy 2024, 26, 579–597. [Google Scholar] [CrossRef]
  11. Bărbulescu, A.; Hosen, K. Cement Industry Pollution and Its Impact on the Environment and Population Health: A Review. Toxics 2025, 13, 587. [Google Scholar] [CrossRef]
  12. Mohamad, N.; Muthusamy, K.; Embong, R.; Kusbiantoro, A.; Hashim, M.H. Environmental impact of cement production and Solutions: A review. Mater. Today Proc. 2021, 48, 741–746. [Google Scholar] [CrossRef]
  13. Bakhtyar, B.; Kacemi, T.; Nawaz, A. International Journal of Energy Economics and Policy A Review on Carbon Emissions in Malaysian Cement Industry. Int. J. Energy Econ. Policy 2017, 7, 282–286. [Google Scholar]
  14. Ghafoor, M.T.; Khan, Q.S.; Qazi, A.U.; Sheikh, M.N.; Hadi, M.N.S. Influence of alkaline activators on the mechanical properties of fly ash based geopolymer concrete cured at ambient temperature. Constr. Build. Mater. 2021, 273, 121752. [Google Scholar] [CrossRef]
  15. Rakhimova, N. Recent Advances in Alternative Cementitious Materials for Nuclear Waste Immobilization: A Review. Sustainability 2023, 15, 689. [Google Scholar] [CrossRef]
  16. Mobili, A.; Telesca, A.; Marroccoli, M.; Tittarelli, F. Calcium sulfoaluminate and alkali-activated fly ash cements as alternative to Portland cement: Study on chemical, physical-mechanical, and durability properties of mortars with the same strength class. Constr. Build. Mater. 2020, 246, 118436. [Google Scholar] [CrossRef]
  17. Alfimova, N.; Levickaya, K.; Elistratkin, M.; Buhtiyarov, I. Supersulfated Cements: A Review Analysis of the Features of Properties Raw Materials Production and Application Prospects. Bull. Belgorod State Technol. Univ. Named After. V. G. Shukhov 2024, 9, 8–24. [Google Scholar] [CrossRef]
  18. Mose, M.P.E.; Perumal, D.B. Latest Advances in Alternative Cementations Binders than Portland cement. IOSR J. Mech. Civ. Eng. 2016, 13, 45–53. [Google Scholar] [CrossRef]
  19. Bernal, S.A.; de Gutierrez, R.M.; Rodríguez, E.D. Alkali-activated materials: Cementing a sustainable future. Ing. Y Compet. 2013, 15, 211–223. [Google Scholar]
  20. Abdullah, M.M.A.; Hussin, K.; Bnhussain, M.; Ismail, K.N.; Ibrahim, W.M.W. Mechanism and Chemical Reaction of Fly Ash Geopolymer Cement—A Review. Asian J. Sci. Res. 2011, 1, 247–253. [Google Scholar]
  21. Davidovits, J. Geopolymer Chemistry and Applications, 5th ed.; Geopolymer Institute: Saint-Quentin, France, 2008; Available online: https://www.researchgate.net/publication/265076752 (accessed on 22 September 2024).
  22. Kumar, A.D.S.; Singh, D.; Reddy, V.S.; Reddy, K.J. Geo-polymerization mechanism and factors affecting it in Metakaolin-slag-fly ash blended concrete. In Proceedings of the E3S Web of Conferences, Chongqing, China, 20–22 November 2020; EDP Sciences: Les Ulis, France, 2020. [Google Scholar] [CrossRef]
  23. Rao, F.; Liu, Q. Geopolymerization and its potential application in mine tailings consolidation: A review. Miner. Process. Extr. Met. Rev. 2015, 36, 399–409. [Google Scholar] [CrossRef]
  24. Zhuang, X.Y.; Chen, L.; Komarneni, S.; Zhou, C.H.; Tong, D.S.; Yang, H.M.; Yu, W.H.; Wang, H. Fly ash-based geopolymer: Clean production, properties and applications. J. Clean. Prod. 2016, 125, 253–267. [Google Scholar] [CrossRef]
  25. Zhang, P.; Gao, Z.; Wang, J.; Guo, J.; Hu, S.; Ling, Y. Properties of fresh and hardened fly ash/slag based geopolymer concrete: A review. J. Clean. Prod. 2020, 270, 122389. [Google Scholar] [CrossRef]
  26. Puertas, F.; González-Fonteboa, B.; González-Taboada, I.; Alonso, M.M.; Torres-Carrasco, M.; Rojo, G.; Martínez-Abella, F. Alkali-activated slag concrete: Fresh and hardened behaviour. Cem. Concr. Compos. 2018, 85, 22–31. [Google Scholar] [CrossRef]
  27. Xie, J.; Chen, W.; Wang, J.; Fang, C.; Zhang, B.; Liu, F. Coupling effects of recycled aggregate and GGBS/metakaolin on physicochemical properties of geopolymer concrete. Constr. Build. Mater. 2019, 226, 345–359. [Google Scholar] [CrossRef]
  28. Pouhet, R.; Cyr, M. Formulation and performance of flash metakaolin geopolymer concretes. Constr. Build. Mater. 2016, 120, 150–160. [Google Scholar] [CrossRef]
  29. Slaty, F.; Khoury, H.; Wastiels, J.; Rahier, H. Characterization of alkali activated kaolinitic clay. Appl. Clay Sci. 2013, 75, 120–125. [Google Scholar] [CrossRef]
  30. Zhang, Z.; Provis, J.L.; Reid, A.; Wang, H. Geopolymer foam concrete: An emerging material for sustainable construction. Constr. Build. Mater. 2014, 56, 113–127. [Google Scholar] [CrossRef]
  31. Singh, N.B.; Middendorf, B. Geopolymers as an alternative to Portland cement: An overview. Constr. Build. Mater. 2020, 237, 117455. [Google Scholar] [CrossRef]
  32. Van Deventer, J.S.J.; Provis, J.L.; Duxson, P. Technical and commercial progress in the adoption of geopolymer cement. Miner. Eng. 2012, 29, 89–104. [Google Scholar] [CrossRef]
  33. Parathi, S.; Nagarajan, P.; Pallikkara, S.A. Ecofriendly geopolymer concrete: A comprehensive review. Clean Technol. Environ. Policy 2021, 23, 1701–1713. [Google Scholar] [CrossRef]
  34. Li, G.; Zhou, C.; Ahmad, W.; Usanova, K.I.; Karelina, M.; Mohamed, A.M.; Khallaf, R. Fly Ash Application as Supplementary Cementitious Material: A Review. Materials 2022, 15, 2664. [Google Scholar] [CrossRef]
  35. Yousuf, A.; Manzoor, S.O.; Youssouf, M.; Malik, Z.A.; Khawaja, K.S. Fly Ash: Production and Utilization in India—An Overview. J. Mater. Environ. Sci. 2020, 11, 911–921. [Google Scholar]
  36. Dwivedi, A.; Jain, M.K. Fly ash-waste management and overview: A Review. Recent Res. Sci. Technol. 2014, 6, 30–35. [Google Scholar]
  37. Gray, D.L. Decrease in Fly Ash Spurring Innovation Within Construction Materials Industry. Nat. Gas Electr. 2019, 35, 23–29. [Google Scholar] [CrossRef]
  38. Saranya, P.; Nagarajan, P.; Shashikala, A.P. Eco-friendly GGBS Concrete: A State-of-The-Art Review. In Proceedings of the IOP Conference Series: Materials Science and Engineering, Hyderabad, India, 1–2 June 2017; Institute of Physics Publishing: Bristol, UK, 2018. [Google Scholar] [CrossRef]
  39. Desta, E.; Jun, Z. A Review on Ground Granulated Blast Slag GGBS in Concrete. In Proceedings of the Eighth International Conference on Advances in Civil and Structural Engineering, CSE, Kuala Lumpur, Malaysia, 4 February 2018; pp. 5–10. [Google Scholar] [CrossRef][Green Version]
  40. Suresh, D.; Nagaraju, K. Ground Granulated Blast Slag (GGBS) In Concrete-A Review. IOSR J. Mech. Civ. Eng. 2015, 12, 76–82. [Google Scholar]
  41. Hoenig, V.; Schall, A.; Sultanov, N.; Papkalla, S.; Ruppert, J. Status and Prospects of Alternative Raw Materials in the European Cement Sector; The European Cement Research Academy (ECRA): Düsseldorf, Germany, 2022. [Google Scholar]
  42. Patil, S.V.; Karikatti, V.B.; Chitawadagi, M. Granulated Blast-Furnace Slag (GGBS) based Geopolymer concrete—Review Concrete—Review. Int. J. Adv. Sci. Eng. 2018, 5, 879. [Google Scholar] [CrossRef]
  43. Hassan, A.; Arif, M.; Shariq, M. Use of geopolymer concrete for a cleaner and sustainable environment—A review of mechanical properties and microstructure. J. Clean. Prod. 2019, 223, 704–728. [Google Scholar] [CrossRef]
  44. Bakera, A.T.; Alexander, M.G. Use of metakaolin as a supplementary cementitious material in concrete, with a focus on durability properties. RILEM Tech. Lett. 2019, 4, 89–102. [Google Scholar] [CrossRef]
  45. Król, M.; Rożek, P. The effect of calcination temperature on metakaolin structure for the synthesis of zeolites. Clay Miner. 2018, 53, 657–663. [Google Scholar] [CrossRef]
  46. Rashad, A.M. Metakaolin as cementitious material: History, scours, production and composition-A comprehensive overview. Constr. Build. Mater. 2013, 41, 303–318. [Google Scholar] [CrossRef]
  47. Kuenzel, C.; Neville, T.P.; Donatello, S.; Vandeperre, L.; Boccaccini, A.R.; Cheeseman, C.R. Influence of metakaolin characteristics on the mechanical properties of geopolymers. Appl. Clay Sci. 2013, 83, 308–314. [Google Scholar] [CrossRef]
  48. Shaikh, F.U.A. Review of mechanical properties of short fibre reinforced geopolymer composites. Constr. Build. Mater. 2013, 43, 37–49. [Google Scholar] [CrossRef]
  49. Patil, S.; Joshi, D.; Mangla, D.; Savvidis, I. Recent development in geopolymer concrete: A review. Mater. Today Proc. 2023. [Google Scholar] [CrossRef]
  50. Venkatesan, G.; Alengaram, U.J.; Ibrahim, S.; Ibrahim, M.S.I. Effect of Fly Ash characteristics, sodium-based alkaline activators, and process variables on the compressive strength of siliceous Fly Ash geopolymers with microstructural properties: A comprehensive review. Constr. Build. Mater. 2024, 437, 136808. [Google Scholar] [CrossRef]
  51. El Alouani, M.; Saufi, H.; Aouan, B.; Bassam, R.; Alehyen, S.; Rachdi, Y.; El Hadki, H.; El Hadki, A.; Mabrouki, J.; Belaaouad, S.; et al. A comprehensive review of synthesis, characterization, and applications of aluminosilicate materials-based geopolymers. Environ. Adv. 2024, 16, 100524. [Google Scholar] [CrossRef]
  52. Zhang, B. Durability of low-carbon geopolymer concrete: A critical review. Sustain. Mater. Technol. 2024, 40, e00882. [Google Scholar] [CrossRef]
  53. Magotra, S.; Jee, A.A. A Review on durability and microstructure of Fly-Ash based geopolymer concrete (FA-GPC). Mater. Today Proc. 2024. [Google Scholar] [CrossRef]
  54. McLellan, B.C.; Williams, R.P.; Lay, J.; Van Riessen, A.; Corder, G.D. Costs and carbon emissions for geopolymer pastes in comparison to ordinary portland cement. J. Clean. Prod. 2011, 19, 1080–1090. [Google Scholar] [CrossRef]
  55. Hajimohammadi, A.; Provis, J.L.; Van Deventer, J.S.J. Effect of alumina release rate on the mechanism of geopolymer gel formation. Chem. Mater. 2010, 22, 5199–5208. [Google Scholar] [CrossRef]
  56. Ma, C.K.; Awang, A.Z.; Omar, W. Structural and material performance of geopolymer concrete: A review. Constr. Build. Mater. 2018, 186, 90–102. [Google Scholar] [CrossRef]
  57. Mabroum, S.; Moukannaa, S.; El Machi, A.; Taha, Y.; Benzaazoua, M.; Hakkou, R. Mine wastes based geopolymers: A critical review. Clean. Eng. Technol. 2020, 1, 100014. [Google Scholar] [CrossRef]
  58. Burciaga-Díaz, O.; Díaz-Guillén, M.R.; Fuentes, A.F.; Escalante-Garcia, J.I. Mortars of alkali-activated blast furnace slag with high aggregate: Binder ratios. Constr. Build. Mater. 2013, 44, 607–614. [Google Scholar] [CrossRef]
  59. Aouan, B.; Alehyen, S.; Fadil, M.; El Alouani, M.; Saufi, H.; Taibi, M. Characteristics, microstructures, and optimization of the geopolymer paste based on three aluminosilicate materials using a mixture design methodology. Constr. Build. Mater. 2023, 384, 131475. [Google Scholar] [CrossRef]
  60. Ryu, G.S.; Lee, Y.B.; Koh, K.T.; Chung, Y.S. The mechanical properties of fly ash-based geopolymer concrete with alkaline activators. Constr. Build. Mater. 2013, 47, 409–418. [Google Scholar] [CrossRef]
  61. Davidovits, J. Properties of geopolymer cements. In Proceedings of the First International Conference on Alkaline Cements and Concretes, Kiev, Ukraine, 11–14 October 1994; pp. 131–149. [Google Scholar]
  62. Davidovits, J. Geopolymer chemistry and sustainable development. The poly (sialate) terminology: A very useful and simple model. In Proceedings of the International Workshop on Geopolymer Cements and Concrete for the Promotion and Understanding of Green-Chemistry, Perth, Australia, 28–29 September 2005; pp. 2181–2278. [Google Scholar]
  63. Nasir, M.; Mahmood, A.H.; Bahraq, A.A. History, recent progress, and future challenges of alkali-activated binders—An overview. Constr. Build. Mater. 2024, 426, 136141. [Google Scholar] [CrossRef]
  64. Yılmaz, A.; Degirmenci, F.N.; Aygörmez, Y. Effect of initial curing conditions on the durability performance of low-calcium fly ash-based geopolymer mortars. Boletín Soc. Española Cerámica Vidr. 2024, 63, 238–254. [Google Scholar] [CrossRef]
  65. Sarmin, S.N.; Welling, J.; Krause, A.; Shalbafan, A. Investigating the Possibility of Geopolymer to Produce Inorganic-Bonded Wood Composites for Multifunctional Construction Material-A Review. Bioresources 2014, 9, 7941–7950. [Google Scholar] [CrossRef]
  66. Abdila, S.R.; Abdullah, M.M.A.B.; Ahmad, R.; Nergis, D.D.B.; Rahim, S.Z.A.; Omar, M.F.; Sandu, A.V.; Vizureanu, P. Syafwandi, Potential of Soil Stabilization Using Ground Granulated Blast Furnace Slag (GGBFS) and Fly Ash via Geopolymerization Method: A Review. Materials 2022, 15, 375. [Google Scholar] [CrossRef]
  67. Yang, K.; White, C.E. Modeling the Formation of Alkali Aluminosilicate Gels at the Mesoscale Using Coarse-Grained Monte Carlo. Langmuir 2016, 32, 11580–11590. [Google Scholar] [CrossRef] [PubMed]
  68. Zhang, M.; Deskins, N.A.; Zhang, G.; Cygan, R.T.; Tao, M. Modeling the Polymerization Process for Geopolymer Synthesis through Reactive Molecular Dynamics Simulations. J. Phys. Chem. C 2018, 122, 6760–6773. [Google Scholar] [CrossRef]
  69. Aleem, A.; Arumairaj, P.D.; Aleem, M.I.A. Pollution free geopolymer concrete with M-Sand. Poll Res. 2013, 32, 139–144. [Google Scholar]
  70. Aleem, M.I.A.; Arumairaj, P.D. Geopolymer Concrete—A Review. Int. J. Eng. Sci. Emerg. Technol. 2012, 1, 118–122. [Google Scholar] [CrossRef]
  71. Palomo, A.; Fernández-Jiménez, A. Alkaline activation, procedure for transforming fly ash into new materials. Part I: Applications. In Proceedings of the World of Coal Ash (WOCA) Conference, Denver, CO, USA, 9–12 May 2011; pp. 1–14. [Google Scholar]
  72. Nawaz, M.; Heitor, A.; Sivakumar, M. Geopolymers in construction—Recent developments. Constr. Build. Mater. 2020, 260, 120472. [Google Scholar] [CrossRef]
  73. Glukhovsky, V.D.; Pashkov, I.A.; Yavorsky, G.A. New building material, in Russian, Bulletin of Technical Information. Gosstrojizdat Kiev 1967, 245, 627. [Google Scholar]
  74. Amran, M.; Al-Fakih, A.; Chu, S.H.; Fediuk, R.; Haruna, S.; Azevedo, A.; Vatin, N. Long-term durability properties of geopolymer concrete: An in-depth review. Case Stud. Constr. Mater. 2021, 15, e00661. [Google Scholar] [CrossRef]
  75. Das, D.; Rout, P.K. Synthesis, Characterization and Properties of Fly Ash Based Geopolymer Materials. J. Mater. Eng. Perform. 2021, 30, 3213–3231. [Google Scholar] [CrossRef]
  76. Fan, F.; Liu, Z.; Xu, G.; Peng, H.; Cai, C.S. Mechanical and thermal properties of fly ash based geopolymers. Constr. Build. Mater. 2018, 160, 66–81. [Google Scholar] [CrossRef]
  77. Lirer, S.; Liguori, B.; Capasso, I.; Flora, A.; Caputo, D. Mechanical and chemical properties of composite materials made of dredged sediments in a fly-ash based geopolymer. J. Environ. Manag. 2017, 191, 1–7. [Google Scholar] [CrossRef] [PubMed]
  78. Shehab, H.K.; Eisa, A.S.; Wahba, A.M. Mechanical properties of fly ash based geopolymer concrete with full and partial cement replacement. Constr. Build. Mater. 2016, 126, 560–565. [Google Scholar] [CrossRef]
  79. Chuah, S.; Duan, W.H.; Pan, Z.; Hunter, E.; Korayem, A.H.; Zhao, X.L.; Collins, F.; Sanjayan, J.G. The properties of fly ash based geopolymer mortars made with dune sand. Mater. Des. 2016, 92, 571–578. [Google Scholar] [CrossRef]
  80. Zerzouri, M.; Bouchenafa, O.; Hamzaoui, R.; Ziyani, L.; Alehyen, S. Physico-chemical and mechanical properties of fly ash based-geopolymer pastes produced from pre-geopolymer powders obtained by mechanosynthesis. Constr. Build. Mater. 2021, 288, 123135. [Google Scholar] [CrossRef]
  81. Shinde, B.H.; Kadam, K.N. Properties of Fly Ash based Geopolymer Mortar. Int. J. Eng. Res. Technol. 2015, 4, 971–974. [Google Scholar]
  82. Wang, Y.; Liu, X.; Zhang, W.; Li, Z.; Zhang, Y.; Li, Y.; Ren, Y. Effects of Si/Al ratio on the efflorescence and properties of fly ash based geopolymer. J. Clean. Prod. 2020, 244, 118852. [Google Scholar] [CrossRef]
  83. Rosyidi, S.A.P.; Rahmad, S.; Yusoff, N.I.M.; Shahrir, A.H.; Ibrahim, A.N.H.; Ismail, N.F.N.; Badri, K.H. Investigation of the chemical, strength, adhesion and morphological properties of fly ash based geopolymer-modified bitumen. Constr. Build. Mater. 2020, 255, 119364. [Google Scholar] [CrossRef]
  84. Hamidi, R.M.; Man, Z.; Azizli, K.A. Concentration of NaOH and the Effect on the Properties of Fly Ash Based Geopolymer. Procedia Eng. 2016, 148, 189–193. [Google Scholar] [CrossRef]
  85. El Alouani, M.; Alehyen, S.; El Achouri, M.; Hajjaji, A.; Ennawaoui, C.; Taibi, M. Influence of the Nature and Rate of Alkaline Activator on the Physicochemical Properties of Fly Ash-Based Geopolymers. Adv. Civ. Eng. 2020, 2020, 8880906. [Google Scholar] [CrossRef]
  86. Okoye, F.N.; Durgaprasad, J.; Singh, N.B. Effect of silica fume on the mechanical properties of fly ash based-geopolymer concrete. Ceram. Int. 2016, 42, 3000–3006. [Google Scholar] [CrossRef]
  87. Huang, Y.; Han, M. The influence of α-Al2O3 addition on microstructure, mechanical and formaldehyde adsorption properties of fly ash-based geopolymer products. J. Hazard. Mater. 2011, 193, 90–94. [Google Scholar] [CrossRef] [PubMed]
  88. Görhan, G.; Kürklü, G. The influence of the NaOH solution on the properties of the fly ash-based geopolymer mortar cured at different temperatures. Compos. B Eng. 2014, 58, 371–377. [Google Scholar] [CrossRef]
  89. Zain, H.; Abdullah, M.M.A.B.; Hussin, K.; Ariffin, N.; Bayuaji, R. Review on Various Types of Geopolymer Materials with the Environmental Impact Assessment. In Proceedings of the MATEC Web of Conferences, Sibiu, Romania, 7–9 June 2017; EDP Sciences: Les Ulis, France, 2017. [Google Scholar] [CrossRef]
  90. Nagajothi, S.; Elavenil, S.; Angalaeswari, S.; Natrayan, L.; Mammo, W.D. Durability Studies on Fly Ash Based Geopolymer Concrete Incorporated with Slag and Alkali Solutions. Adv. Civ. Eng. 2022, 2022, 7196446. [Google Scholar] [CrossRef]
  91. Rihan, M.A.M.; Alahmari, T.S.; Onchiri, R.O.; Gathimba, N.; Sabuni, B. Impact of Alkaline Concentration on the Mechanical Properties of Geopolymer Concrete Made up of Fly Ash and Sugarcane Bagasse Ash. Sustainability 2024, 16, 2841. [Google Scholar] [CrossRef]
  92. Bajpai, R.; Choudhary, K.; Srivastava, A.; Sangwan, K.S.; Singh, M. Environmental impact assessment of fly ash and silica fume based geopolymer concrete. J. Clean. Prod. 2020, 254, 120147. [Google Scholar] [CrossRef]
  93. Bellum, R.R.; Venkatesh, C.; Madduru, S.R.C. Influence of red mud on performance enhancement of fly ash-based geopolymer concrete. Innov. Infrastruct. Solut. 2021, 6, 215. [Google Scholar] [CrossRef]
  94. Kumar, A.; Kumar, S. Development of paving blocks from synergistic use of red mud and fly ash using geopolymerization. Constr. Build. Mater. 2013, 38, 865–871. [Google Scholar] [CrossRef]
  95. Abdullah, A.; Hussin, K.; Abdullah, M.M.A.B.; Yahya, Z.; Sochacki, W.; Razak, R.A.; Błoch, K.; Fansuri, H. The effects of various concentrations of naoh on the inter-particle gelation of a fly ash geopolymer aggregate. Materials 2021, 14, 1111. [Google Scholar] [CrossRef]
  96. Wulandari, K.D.; Ekaputri, J.J.; Triwulan; Kurniawan, S.B.; Primaningtyas, W.E.; Abdullah, S.R.S.; Ismail, I.; Imron, M.F. Effect of microbes addition on the properties and surface morphology of fly ash-based geopolymer paste. J. Build. Eng. 2021, 33, 101596. [Google Scholar] [CrossRef]
  97. Ouyang, X.; Ma, Y.; Liu, Z.; Liang, J.; Ye, G. Effect of the sodium silicate modulus and slag content on fresh and hardened properties of alkali-activated fly ash/slag. Minerals 2020, 10, 15. [Google Scholar] [CrossRef]
  98. Sivasakthi, M.; Jeyalakshmi, R.; Rajamane, N.P. Fly ash geopolymer mortar: Impact of the substitution of river sand by copper slag as a fine aggregate on its thermal resistance properties. J. Clean. Prod. 2021, 279, 123766. [Google Scholar] [CrossRef]
  99. Błaszczyński, T.Z.; Król, M.R. Geopolymers in Construction/Zastosowanie Geopolimerów W Budownictwie. Civ. Environ. Eng. Rep. 2023, 16, 25–40. [Google Scholar] [CrossRef][Green Version]
  100. Fernández-Jiménez, A.; Palomo, A.; Sobrados, I.; Sanz, J. The role played by the reactive alumina content in the alkaline activation of fly ashes. Microporous Mesoporous Mater. 2006, 91, 111–119. [Google Scholar] [CrossRef]
  101. Soutsos, M.; Boyle, A.P.; Vinai, R.; Hadjierakleous, A.; Barnett, S.J. Factors influencing the compressive strength of fly ash based geopolymers. Constr. Build. Mater. 2016, 110, 355–368. [Google Scholar] [CrossRef]
  102. Diaz, E.I.; Allouche, E.N.; Eklund, S. Factors affecting the suitability of fly ash as source material for geopolymers. Fuel 2010, 89, 992–996. [Google Scholar] [CrossRef]
  103. Alehyen, S.; Achouri, M.E.L.; Taibi, M. Characterization, microstructure, and properties of fly ash-based geopolymer. J. Mater. Environ. Sci. 2017, 8, 1783–1796. [Google Scholar]
  104. Yazdi, M.A.; Liebscher, M.; Hempel, S.; Yang, J.; Mechtcherine, V. Correlation of microstructural and mechanical properties of geopolymers produced from fly ash and slag at room temperature. Constr. Build. Mater. 2018, 191, 330–341. [Google Scholar] [CrossRef]
  105. Nagalia, G.; Park, Y.; Abolmaali, A.; Aswath, P. Compressive Strength and Microstructural Properties of Fly Ash–Based Geopolymer Concrete. J. Mater. Civ. Eng. 2016, 28, 1656. [Google Scholar] [CrossRef]
  106. Bohra, V.K.J.; Nerella, R.; Madduru, S.R.C.; Rohith, P. Microstructural characterization of fly ash based geopolymer. Mater. Today Proc. 2020, 27, 1625–1629. [Google Scholar] [CrossRef]
  107. Alahmad, J.; BiBi, A.; Al-Ghouti, M.A. Application of TiO2-loaded fly ash-based geopolymer in adsorption of methylene blue from water: Waste-to-value approach. Groundw. Sustain. Dev. 2024, 25, 101138. [Google Scholar] [CrossRef]
  108. Xu, H.; Gong, W.; Syltebo, L.; Izzo, K.; Lutze, W.; Pegg, I.L. Effect of blast furnace slag grades on fly ash based geopolymer waste forms. Fuel 2014, 133, 332–340. [Google Scholar] [CrossRef]
  109. Aziz, I.H.; Abdullah, M.M.A.B.; Salleh, M.A.A.M.; Azimi, E.A.; Chaiprapa, J.; Sandu, A.V. Strength development of solely ground granulated blast furnace slag geopolymers. Constr. Build. Mater. 2020, 250, 118720. [Google Scholar] [CrossRef]
  110. Kumar, S.; Mucsi, G.; Kristály, F.; Pekker, P. Mechanical activation of fly ash and its influence on micro and nano-structural behaviour of resulting geopolymers. Adv. Powder Technol. 2017, 28, 805–813. [Google Scholar] [CrossRef]
  111. Rożek, P.; Król, M.; Mozgawa, W. Spectroscopic studies of fly ash-based geopolymers. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2018, 198, 283–289. [Google Scholar] [CrossRef]
  112. Nath, S.K. Geopolymerization behavior of ferrochrome slag and fly ash blends. Constr. Build. Mater. 2018, 181, 487–494. [Google Scholar] [CrossRef]
  113. Mahfoud, E.; Maherzi, W.; Ndiaye, K.; Benzerzour, M.; Aggoun, S.; Abriak, N.E. Mechanical and microstructural properties of just add water geopolymer cement comprised of Thermo-Mechanicalsynthesis Sediments-Fly ash mix. Constr. Build. Mater. 2023, 400, 132626. [Google Scholar] [CrossRef]
  114. Nath, S.K.; Maitra, S.; Mukherjee, S.; Kumar, S. Microstructural and morphological evolution of fly ash based geopolymers. Constr. Build. Mater. 2016, 111, 758–765. [Google Scholar] [CrossRef]
  115. Luhar, S.; Chaudhary, S.; Luhar, I. Thermal resistance of fly ash based rubberized geopolymer concrete. J. Build. Eng. 2018, 19, 420–428. [Google Scholar] [CrossRef]
  116. Rosas, C.A.; Arredondo-Rea, S.P.; Gómez-Soberón, J.M.; Almaral-Sánchez, J.L.; Rosas-Casarez, C.A.; Arredondo-Rea, S.P.; Gómez-Soberón, J.M.; Alamaral-Sánchez, J.L.; Corral-Higuera, R.; Chinchillas-Chinchillas, M.J.; et al. Experimental Study of XRD, FTIR and TGA Techniques in Geopolymeric Materials, 2014. Available online: https://www.researchgate.net/publication/274079395 (accessed on 30 November 2024).
  117. Li, S.; Huang, X.; Muhammad, F.; Yu, L.; Xia, M.; Zhao, J.; Jiao, B.; Shiau, Y.; Li, D. Waste solidification/stabilization of lead-zinc slag by utilizing fly ash based geopolymers. RSC Adv. 2018, 8, 32956–32965. [Google Scholar] [CrossRef]
  118. Öz, H.Ö.; Yücel, H.E.; Güneş, M.; Köker, T.Ş. Fly-ash-based geopolymer composites incorporating cold-bonded lightweight fly ash aggregates. Constr. Build. Mater. 2021, 272, 121963. [Google Scholar] [CrossRef]
  119. Caballero, L.R.; Paiva, M.D.D.M.; Fairbairn, E.D.M.R.; Filho, R.D.T. Thermal, mechanical and microstructural analysis of metakaolin based geopolymers. Mater. Res. 2019, 22, e20180716. [Google Scholar] [CrossRef]
  120. Catauro, M.; Tranquillo, E.; Barrino, F.; Dal Poggetto, G.; Blanco, I.; Cicala, G.; Ognibene, G.; Recca, G. Mechanical and thermal properties of fly ash-filled geopolymers. J. Therm. Anal. Calorim. 2019, 138, 3267–3276. [Google Scholar] [CrossRef]
  121. Zhou, W.; Yan, C.; Duan, P.; Liu, Y.; Zhang, Z.; Qiu, X.; Li, D. A comparative study of high- and low-Al2O3 fly ash based-geopolymers: The role of mix proportion factors and curing temperature. Mater. Des. 2016, 95, 63–74. [Google Scholar] [CrossRef]
  122. Du, W.; Tang, N.; Sun, Y.; Nie, N.; Zhang, R.; Wang, K. Adhesion performance and enhancement mechanism of FA/GGBFS based geopolymer modified bitumen and acidic aggregate. Case Stud. Constr. Mater. 2024, 21, e03850. [Google Scholar] [CrossRef]
  123. Wang, T.; Fan, X.; Gao, C. Strength, pore characteristics, and characterization of fly ash-slag-based geopolymer mortar modified with silica fume. Structures 2024, 69, 107525. [Google Scholar] [CrossRef]
  124. Nath, S.K.; Kumar, S. Evaluation of the suitability of ground granulated silico-manganese slag in Portland slag cement. Constr. Build. Mater. 2016, 125, 127–134. [Google Scholar] [CrossRef]
  125. Gao, X.; Yu, Q.L.; Brouwers, H.J.H. Reaction kinetics, gel character and strength of ambient temperature cured alkali activated slag-fly ash blends. Constr. Build. Mater. 2015, 80, 105–115. [Google Scholar] [CrossRef]
  126. Li, Z.; Zhang, J.; Lei, Z.; Gao, M.; Sun, J.; Tong, L.; Chen, S.; Wang, Y. Designing low-carbon fly ash based geopolymer with red mud and blast furnace slag wastes: Performance, microstructure and mechanism. J. Environ. Manag. 2024, 354, 120362. [Google Scholar] [CrossRef]
  127. Zhang, M.; He, M.; Pan, Z. Inhibition of efflorescence for fly ash-slag-steel slag based geopolymer: Pore network optimization and free alkali stabilization. Ceram. Int. 2024, 50, 48538–48550. [Google Scholar] [CrossRef]
  128. Xu, X.; Bao, S.; Zhang, Y.; Ping, Y. Sustainable enhancement of fly ash-based geopolymers: Impact of Alkali thermal activation and particle size on green production. Process Saf. Environ. Prot. 2024, 191, 478–489. [Google Scholar] [CrossRef]
  129. Li, H.; Gao, P.; Xu, F.; Sun, T.; Zhou, Y.; Zhu, J.; Peng, C.; Lin, J. Effect of fine aggregate particle characteristics on mechanical properties of fly ash-based geopolymer mortar. Minerals 2021, 11, 897. [Google Scholar] [CrossRef]
  130. Nagajothi, S.; Elavenil, S. Effect of GGBS Addition on Reactivity and Microstructure Properties of Ambient Cured Fly Ash Based Geopolymer Concrete. Silicon 2021, 13, 507–516. [Google Scholar] [CrossRef]
  131. Yu, G.; Jia, Y. Microstructure and Mechanical Properties of Fly Ash-Based Geopolymer Cementitious Composites. Minerals 2022, 12, 853. [Google Scholar] [CrossRef]
  132. Lv, Q.; Yu, J.; Ji, F.; Gu, L.; Chen, Y.; Shan, X. Mechanical Property and Microstructure of Fly Ash-Based Geopolymer Activated by Sodium Silicate. KSCE J. Civ. Eng. 2021, 25, 1765–1777. [Google Scholar] [CrossRef]
  133. Razak, S.N.A.; Shafiq, N.; Guillaumat, L.; Farhan, S.A.; Lohana, V.K. Fire-Exposed Fly-Ash-Based Geopolymer Concrete: Effects of Burning Temperature on Mechanical and Microstructural Properties. Materials 2022, 15, 1884. [Google Scholar] [CrossRef]
  134. Hager, I.; Sitarz, M.; Mróz, K. Fly-ash based geopolymer mortar for high-temperature application—Effect of slag addition. J. Clean. Prod. 2021, 316, 128168. [Google Scholar] [CrossRef]
  135. Bellum, R.R.; Muniraj, K.; Madduru, S.R.C. Influence of slag on mechanical and durability properties of fly ash-based geopolymer concrete. J. Korean Ceram. Soc. 2020, 57, 530–545. [Google Scholar] [CrossRef]
  136. Do, Q.M.; Ngo, P.M.; Nguyen, H.T. Characteristics of a fly ash-based geopolymer cured in microwave oven. Key Eng. Mater. 2020, 850, 63–69. [Google Scholar] [CrossRef]
  137. Kuenzel, C.; Ranjbar, N. Dissolution mechanism of fly ash to quantify the reactive aluminosilicates in geopolymerisation. Resour. Conserv. Recycl. 2019, 150, 104421. [Google Scholar] [CrossRef]
  138. Nath, S.K.; Kumar, S. Role of particle fineness on engineering properties and microstructure of fly ash derived geopolymer. Constr. Build. Mater. 2020, 233, 117294. [Google Scholar] [CrossRef]
  139. Cai, J.; Pan, J.; Li, X.; Tan, J.; Li, J. Electrical resistivity of fly ash and metakaolin based geopolymers. Constr. Build. Mater. 2020, 234, 117868. [Google Scholar] [CrossRef]
  140. Abdullah, M.M.A.B.; Hussin, K.; Binhussain, M.; Nizar, I.K.; Razak, R.A.; Zarina, Y. Microstructure study on optimization of high strength fly ash based geopolymer. Adv. Mater. Res. 2012, 476, 2173–2180. [Google Scholar] [CrossRef]
  141. Zhang, Y.J.; Li, H.H.; Wang, Y.C.; Xu, D.L. Geopolymer microstructure and hydration mechanism of alkali-activated fly ash-based geopolymer. Adv. Mater. Res. 2012, 374, 1481–1484. [Google Scholar] [CrossRef]
  142. Rawat, R.; Pasla, D. Assessment of mechanical and durability properties of FA-GGBS based lightweight geopolymer concrete. Constr. Build. Mater. 2024, 426, 135984. [Google Scholar] [CrossRef]
  143. Li, J.; Liu, X.; Cai, C.; Li, F.; Jin, H.; Zhang, S.; Sheng, N.; Wang, T.; Ngo, T. Microstructure, strength development mechanism and CO2 emission assessments of molybdenum tailings collaborative fly ash geopolymers. Case Stud. Constr. Mater. 2024, 21, e03608. [Google Scholar] [CrossRef]
  144. Li, Z.; Chen, R.; Zhang, L. Utilization of chitosan biopolymer to enhance fly ash-based geopolymer. J. Mater. Sci. 2013, 48, 7986–7993. [Google Scholar] [CrossRef]
  145. Khan, M.I.; Azizli, K.; Sufian, S.; Man, Z.; Siyal, A.A.; Ullah, H. Determination of anisotropy and multimorphology in fly ash based geopolymers. In Proceedings of the AIP Conference Proceeding, Tronoh, Malaysia, 10–12 December 2014; American Institute of Physics Inc.: College Park, MD, USA, 2015. [Google Scholar] [CrossRef]
  146. Çeli, S.; Öztürk, Z.B.; Atabey, İ.İ. High-temperature resistance of ceramic sanitaryware waste and fly ash-based geopolymer and hybrid geopolymer mortars produced at ambient curing conditions. Constr. Build. Mater. 2024, 446, 137990. [Google Scholar] [CrossRef]
  147. Bellum, R.R.; Al Khazaleh, M.; Pilla, R.K.; Choudhary, S.; Venkatesh, C. Effect of slag on strength, durability and microstructural characteristics of fly ash-based geopolymer concrete. J. Build. Pathol. Rehabil. 2022, 7, 25. [Google Scholar] [CrossRef]
  148. Dinh, H.L.; Liu, J.; Doh, J.H.; Ong, D.E.L. Influence of Si/Al molar ratio and ca content on the performance of fly ash-based geopolymer incorporating waste glass and GGBFS. Constr. Build. Mater. 2024, 411, 134741. [Google Scholar] [CrossRef]
  149. Liu, J.; Doh, J.H.; Ong, D.E.L.; Dinh, H.L.; Podolsky, Z.; Zi, G. Investigation on red mud and fly ash-based geopolymer: Quantification of reactive aluminosilicate and derivation of effective Si/Al molar ratio. J. Build. Eng. 2023, 71, 106559. [Google Scholar] [CrossRef]
  150. Ling, Y.; Wang, K.; Li, W.; Shi, G.; Lu, P. Effect of slag on the mechanical properties and bond strength of fly ash-based engineered geopolymer composites. Compos. B Eng. 2019, 164, 747–757. [Google Scholar] [CrossRef]
  151. Zhao, D.; Li, K.; Li, Y.; Chen, X.; Fan, J.; Zhu, J. Self-healing behaviour of fly ash/slag-based engineered geopolymer composites under external alkaline environments. Case Stud. Constr. Mater. 2024, 21, e03361. [Google Scholar] [CrossRef]
  152. Li, F.; Liu, L.; Yang, Z.; Li, S. Physical and mechanical properties and micro characteristics of fly ash-based geopolymer paste incorporated with waste Granulated Blast Furnace Slag (GBFS) and functionalized Multi-Walled Carbon Nanotubes (MWCNTs). J. Hazard. Mater. 2021, 401, 123339. [Google Scholar] [CrossRef] [PubMed]
  153. Li, R.; Wu, G.; Jiang, L.; Sun, D. Characterization of multi-scale porous structure of fly ash/phosphate geopolymer hollow sphere structures: From submillimeter to nano-scale. Micron 2014, 68, 54–58. [Google Scholar] [CrossRef]
  154. Ranjbar, N.; Mehrali, M.; Mehrali, M.; Alengaram, U.J.; Jumaat, M.Z. Graphene nanoplatelet-fly ash based geopolymer composites. Cem. Concr. Res. 2015, 76, 222–231. [Google Scholar] [CrossRef]
  155. Yang, Z.; Hao, H.; Lu, Y.; Li, S. Effects of Cl- erosion conditions on Cl- binding properties of fly-ash-slag-based geopolymer. J. Build. Eng. 2023, 67, 105907. [Google Scholar] [CrossRef]
  156. Sivasakthi, M.; Jeyalakshmi, R.; Rajamane, N.P.; Jose, R. Thermal and structural micro analysis of micro silica blended fly ash based geopolymer composites. J. Non Cryst. Solids 2018, 499, 117–130. [Google Scholar] [CrossRef]
  157. Ranjbar, N.; Mehrali, M.; Alengaram, U.J.; Metselaar, H.S.C.; Jumaat, M.Z. Compressive strength and microstructural analysis of fly ash/palm oil fuel ash based geopolymer mortar under elevated temperatures. Constr. Build. Mater. 2014, 65, 114–121. [Google Scholar] [CrossRef]
  158. Hosan, A.; Haque, S.; Shaikh, F. Compressive behaviour of sodium and potassium activators synthetized fly ash geopolymer at elevated temperatures: A comparative study. J. Build. Eng. 2016, 8, 123–130. [Google Scholar] [CrossRef]
  159. Wang, Y.C.; Zhang, Y.J.; Xu, D.L. Novel activator for synthesis of fly ash based geopolymer. Mater. Res. Innov. 2014, 18, S2238–S2242. [Google Scholar] [CrossRef]
  160. Cui, Y.; Wang, D.; Wang, Y.; Sun, R.; Rui, Y. Effects of the n(H2O: Na2Oeq) ratio on the geopolymerization process and microstructures of fly ash-based geopolymers. J. Non Cryst. Solids 2019, 511, 19–28. [Google Scholar] [CrossRef]
  161. Alomayri, T.; Shaikh, F.U.A.; Low, I.M. Mechanical and thermal properties of ambient cured cotton fabric-reinforced fly ash-based geopolymer composites. Ceram. Int. 2014, 40, 14019–14028. [Google Scholar] [CrossRef]
  162. Cong, P.; Mei, L. Using silica fume for improvement of fly ash/slag based geopolymer activated with calcium carbide residue and gypsum. Constr. Build. Mater. 2021, 275, 122171. [Google Scholar] [CrossRef]
  163. Jiang, X.; Zhang, Y.; Xiao, R.; Polaczyk, P.; Zhang, M.; Hu, W.; Bai, Y.; Huang, B. A comparative study on geopolymers synthesized by different classes of fly ash after exposure to elevated temperatures. J. Clean. Prod. 2020, 270, 122500. [Google Scholar] [CrossRef]
  164. Alghamdi, H.; Neithalath, N. Novel synthesis of lightweight geopolymer matrices from fly ash through carbonate-based activation. Mater. Today Commun. 2018, 17, 266–277. [Google Scholar] [CrossRef]
  165. Shi, S.; Li, H.; Zhou, Q.; Zhang, H.; Basheer, P.A.M.; Bai, Y. Alkali-activated fly ash cured with pulsed microwave and thermal oven: A comparison of reaction products, microstructure and compressive strength. Cem. Concr. Res. 2023, 166, 107104. [Google Scholar] [CrossRef]
  166. Mahfoud, E.; Ndiaye, K.; Maherzi, W.; Aggoun, S.; Abriak, N.-E.; Benzerzour, M. Carbonation in just add water geopolymer based on fly ash and dredged sediments mix. Constr. Build. Mater. 2024, 452, 138959. [Google Scholar] [CrossRef]
  167. Öz, H.Ö.; Güneş, M.; Yücel, H.E. Rheological and microstructural properties of FA+GGBFS-based engineered geopolymer composites (EGCs) capable of comparing with M45-ECC as mechanical performance. J. Build. Eng. 2023, 65, 105792. [Google Scholar] [CrossRef]
  168. Jiang, X.; Xiao, R.; Zhang, M.; Hu, W.; Bai, Y.; Huang, B. A laboratory investigation of steel to fly ash-based geopolymer paste bonding behavior after exposure to elevated temperatures. Constr. Build. Mater. 2020, 254, 119267. [Google Scholar] [CrossRef]
  169. Rickard, W.D.A.; Van Riessen, A.; Walls, P. Thermal character of geopolymers synthesized from class Fly ash containing high concentrations of iron and α-quartz. Int. J. Appl. Ceram. Technol. 2010, 7, 81–88. [Google Scholar] [CrossRef]
  170. Alexander, A.E.; Shashikala, A.P. Studies on the microstructure and durability characteristics of ambient cured FA-GGBS based geopolymer mortar. Constr. Build. Mater. 2022, 347, 128538. [Google Scholar] [CrossRef]
  171. Öz, H.Ö.; Doğan-Sağlamtimur, N.; Bilgil, A.; Tamer, A.; Günaydin, K. Process development of fly ash-based geopolymer mortars in view of the mechanical characteristics. Materials 2021, 14, 2935. [Google Scholar] [CrossRef] [PubMed]
  172. Aziz, I.H.; Al Bakri Abdullah, M.M.; Heah, C.Y.; Liew, Y.M. Behaviour changes of ground granulated blast furnace slag geopolymers at high temperature. Adv. Cem. Res. 2020, 32, 465–475. [Google Scholar] [CrossRef]
  173. Sultan, M.E.; Abo-El-Enein, S.A.; Sayed, A.Z.; EL-Sokkary, T.M.; Hammad, H.A. Incorporation of cement bypass flue dust in fly ash and blast furnace slag-based geopolymer. Case Stud. Constr. Mater. 2018, 8, 315–322. [Google Scholar] [CrossRef]
  174. Zhao, J.; Wang, K.; Wang, S.; Wang, Z.; Yang, Z.; Shumuye, E.D.; Gong, X. Effect of elevated temperature on mechanical properties of high-volume fly ash-based geopolymer concrete, mortar and paste cured at room temperature. Polymers 2021, 13, 1473. [Google Scholar] [CrossRef] [PubMed]
  175. Lahoti, M.; Wijaya, S.F.; Tan, K.H.; Yang, E.H. Tailoring sodium-based fly ash geopolymers with variegated thermal performance. Cem. Concr. Compos. 2020, 107, 103507. [Google Scholar] [CrossRef]
  176. Assi, L.N.; Deaver, E.E.; Ziehl, P. Effect of source and particle size distribution on the mechanical and microstructural properties of fly Ash-Based geopolymer concrete. Constr. Build. Mater. 2018, 167, 372–380. [Google Scholar] [CrossRef]
  177. Huang, Y.; Gong, L.; Shi, L.; Cao, W.; Pan, Y.; Cheng, X. Experimental investigation on the influencing factors of preparing porous fly ash-based geopolymer for insulation material. Energy Build. 2018, 168, 9–18. [Google Scholar] [CrossRef]
  178. Zhang, T.; Yang, Z.; Zhang, D.; Yang, Q. Properties of Gangue Powder Modified Fly Ash-Based Geopolymer. Materials 2023, 16, 5719. [Google Scholar] [CrossRef]
  179. Kumar, E.M.; Ramamurthy, K. Influence of production on the strength, density and water absorption of aerated geopolymer paste and mortar using Class F fly ash. Constr. Build. Mater. 2017, 156, 1137–1149. [Google Scholar] [CrossRef]
  180. Sun, Q.; Zhao, S.; Zhao, X.; Song, Y.; Ban, X.; Zhang, N. Influence of different grinding degrees of fly ash on properties and reaction degrees of geopolymers. PLoS ONE 2023, 18, e0282927. [Google Scholar] [CrossRef] [PubMed]
  181. Xu, F.; Gu, G.; Zhang, W.; Wang, H.; Huang, X.; Zhu, J. Pore structure analysis and properties evaluations of fly ash-based geopolymer foams by chemical foaming method. Ceram. Int. 2018, 44, 19989–19997. [Google Scholar] [CrossRef]
  182. Cui, Y.; Wang, D. Effects of Water on Pore Structure and Thermal Conductivity of Fly Ash-Based Foam Geopolymers. Adv. Mater. Sci. Eng. 2019, 2019, 3202794. [Google Scholar] [CrossRef]
  183. Kroehong, W.; Jaturapitakkul, C.; Pothisiri, T.; Chindaprasirt, P. Effect of Oil Palm Fiber Content on the Physical and Mechanical Properties and Microstructure of High-Calcium Fly Ash Geopolymer Paste. Arab. J. Sci. Eng. 2018, 43, 5215–5224. [Google Scholar] [CrossRef]
  184. Hartono, E.; Diana, W.; Nugraha, Y.; Afzalurrahman, M. View of Compaction Properties of the Fly-Ash Based Geopolymer on Silt Soil. Int. J. Geomate 2024, 26, 19–26. [Google Scholar] [CrossRef]
  185. Tenepalli, J.S.; Neeraja, D. Properties of class F fly ash based geopolymer mortar produced with alkaline water. J. Build. Eng. 2018, 19, 42–48. [Google Scholar] [CrossRef]
  186. Haq, E.U.; Padmanabhan, S.K.; Licciulli, A. Synthesis and characteristics of fly ash and bottom ash based geopolymers-A comparative study. Ceram. Int. 2014, 40, 2965–2971. [Google Scholar] [CrossRef]
  187. Ridtirud, C.; Chindaprasirt, P. Properties of lightweight aerated geopolymer synthesis from high-calcium fly ash and aluminium powder. Int. J. GEOMATE 2019, 16, 67–75. [Google Scholar] [CrossRef]
  188. Kamhangrittirong, P.; Suwanvitaya, P.; Witayakul, W.; Suwanvitaya, P.; Chindaprasirt, P. Factors influence on shrinkage of high calcium fly ash geopolymer paste. Adv. Mater. Res. 2013, 610, 2275–2281. [Google Scholar] [CrossRef]
  189. Deraman, L.M.; Abdullah, M.M.A.B.; Ming, L.Y.; Hussin, K. Density and morphology studies on bottom ash and fly ash geopolymer brick. In Proceedings of the AIP Conference Proceedings, Kaohsiung City, Taiwan, 8–9 December 2016; American Institute of Physics Inc.: College Park, MD, USA, 2017. [Google Scholar] [CrossRef]
  190. Azevedo, A.G.; Strecker, K. Production of Fly Ash Based Geopolymers Using Activator Solution with Different Na2O and Na2SiO3 Compositions. Rev. Tecnológica Mar. 2017, 26, 95–102. [Google Scholar] [CrossRef]
  191. Hui-Teng, N.; Cheng-Yong, H.; Abdullah, M.M.A.B.; Yong-Sing, N.; Bayuaji, R. Study of fly ash geopolymer and fly ash/slag geopolymer in term of physical and mechanical properties. Eur. J. Mater. Sci. Eng. 2020, 5, 187–198. [Google Scholar] [CrossRef]
  192. Bijeljić, J.; Ristić, N.; Grdić, D.; Pavlović, M. Possibilities of Biomass Wood Ash Usage in Geopolymer Mixtures. Teh. Vjesn. 2023, 30, 52–60. [Google Scholar] [CrossRef]
  193. Hilal, A.A.; Nader, A.S.; Kurji, B.M. Producing Sustainable Lightweight Geopolymer Concrete Using Waste Materials. Civ. Environ. Eng. 2024, 20, 1120–1128. [Google Scholar] [CrossRef]
  194. Ibrahim, W.M.W.; AL Mustafa Bakri Abdullah, A.M.; Victor Sandu, K.; Hussin, I.; Gabriel Sandu, K.; Nizar Ismail, A.; Abdul Kadir, M.; Binhussain, P.; Raja, B.; Pahat Johor, K.A. Processing and Characterization of Fly Ash-Based Geopolymer Bricks. Rev. Chim 2014, 65, 1340–1345. [Google Scholar]
  195. Nurgesang, F.A.; Wattanasiriwech, S.; Wattanasiriwech, D.; Aungkavattana, P. Mechanical Physical Properties of Fly Ash Geopolymer-Mullite Composites Suranaree. J. Sci. Technol. 2016, 23, 45–52. [Google Scholar]
  196. Ibrahim, W.M.W.; Hussin, K.; Abdullah, M.M.A.; Kadir, A.A.; Deraman, L.M. Effects of sodium hydroxide (NaOH) solution concentration on fly ash-based lightweight geopolymer. In Proceedings of the AIP Conference Proceedings, Krabi, Thailand, 29–30 April 2017; American Institute of Physics Inc.: College Park, MD, USA, 2017. [Google Scholar] [CrossRef]
  197. Abdullah, S.F.A.; Yun-Ming, L.; Al Bakri, M.M.; Cheng-Yong, H.; Zulkifly, K.; Hussin, K. Effect of Alkali Concentration on Fly Ash Geopolymers. In Proceedings of the IOP Conference Series: Materials Science and Engineering, Bali, Indonesia, 25–26 October 2017; Institute of Physics Publishing: Bristol, UK, 2018. [Google Scholar] [CrossRef]
  198. Nematollahi, B.; Sanjayan, J.; Chai, J.X.H.; Lu, T.M. Properties of fresh and hardened glass fiber reinforced fly ash based geopolymer concrete. Key Eng. Mater. 2014, 594, 629–633. [Google Scholar] [CrossRef]
  199. Razak, R.A.; Abdullah, M.M.A.B.; Yahya, Z.; Jian, A.Z.; Nasri, A. Performance of fly ash based geopolymer incorporating palm kernel shell for lightweight concrete. In Proceedings of the AIP Conference Proceedings, Krabi, Thailand, 29–30 April 2017; American Institute of Physics Inc.: College Park, MD, USA, 2017. [Google Scholar] [CrossRef]
  200. Tho-In, T.; Sata, V.; Chindaprasirt, P.; Jaturapitakkul, C. Pervious high-calcium fly ash geopolymer concrete. Constr. Build. Mater. 2012, 30, 366–371. [Google Scholar] [CrossRef]
  201. Wongsa, A.; Zaetang, Y.; Sata, V.; Chindaprasirt, P. Properties of lightweight fly ash geopolymer concrete containing bottom ash as aggregates. Constr. Build. Mater. 2016, 111, 637–643. [Google Scholar] [CrossRef]
  202. Ali, I.M.; Naje, A.S.; Al-Zubaidi, H.A.M.; Al-Kateeb, R.T. Performance evaluation of fly ash-based geopolymer concrete incorporating nano slag. Glob. Nest J. 2019, 21, 70–75. [Google Scholar] [CrossRef]
  203. Wongkeo, W.; Seekaew, S.; Kaewrahan, O. Properties of High Calcium Fly Ash Geopolymer Lightweight Concrete. Mater. Today Proc. 2019, 17, 1423–1430. [Google Scholar] [CrossRef]
  204. Priyadharshini, P.; Ramamurthy, K.; Robinson, R.G. Excavated soil waste as fine aggregate in fly ash based geopolymer mortar. Appl. Clay Sci. 2017, 146, 81–91. [Google Scholar] [CrossRef]
  205. Al Ghafri, E.; Al Tamimi, N.; El-Hassan, H.; Maraqa, M.A.; Hamouda, M. Synthesis and multi-objective optimization of fly ash-slag geopolymer sorbents for heavy metal removal using a hybrid Taguchi-TOPSIS approach. Environ. Technol. Innov. 2024, 35, 103721. [Google Scholar] [CrossRef]
  206. Nalewajko, M.; Bołtryk, M. Use of Alkaline-Activated Energy Waste Raw Materials in Geopolymer Concrete. Materials 2024, 17, 2253. [Google Scholar] [CrossRef] [PubMed]
  207. Naenudon, S.; Vilaivong, A.; Zaetang, Y.; Tangchirapat, W.; Wongsa, A.; Sata, V.; Chindaprasirt, P. High flexural strength lightweight fly ash geopolymer mortar containing waste fiber cement. Case Stud. Constr. Mater. 2022, 16, e01121. [Google Scholar] [CrossRef]
  208. Cui, Y.; Wang, L.; Liu, J.; Liu, R.; Pang, B. Impact of particle size of fly ash on the early compressive strength of concrete: Experimental investigation and modelling. Constr. Build. Mater. 2022, 323, 126444. [Google Scholar] [CrossRef]
  209. Li, Y.; Dong, Y.; El-Naggar, M.R.; Wang, F.; Zhao, Y. The Influence of Particle Size and Calcium Content on Performance Characteristics of Metakaolin- and Fly-Ash-Based Geopolymer Gels. Gels 2024, 10, 639. [Google Scholar] [CrossRef]
  210. Xiong, L.; Wan, Z.; Zhang, Y.; Wang, F.; Wang, J.; Kang, Y. Fly ash particle size effect on pore structure and strength of fly ash foamed geopolymer. Adv. Polym. Technol. 2019, 2019, 1098027. [Google Scholar] [CrossRef]
  211. Wang, Y.; Zheng, T.; Zheng, X.; Liu, Y.; Darkwa, J.; Zhou, G. Thermo-mechanical and moisture absorption properties of fly ash-based lightweight geopolymer concrete reinforced by polypropylene fibers. Constr. Build. Mater. 2020, 251, 118960. [Google Scholar] [CrossRef]
  212. Irum, S.; Shabbir, F. Performance of fly ash/GGBFS based geopolymer concrete with recycled fine and coarse aggregates at hot and ambient curing. J. Build. Eng. 2024, 95, 110148. [Google Scholar] [CrossRef]
  213. Malkawi, A.B.; Nuruddin, M.F.; Fauzi, A.; Almattarneh, H.; Mohammed, B.S. Effects of Alkaline Solution on Properties of the HCFA Geopolymer Mortars. Procedia Eng. 2016, 148, 710–717. [Google Scholar] [CrossRef]
  214. Zhao, X.; Wang, H.; Jiang, L.; Meng, L.; Zhou, B.; Zhang, J. Long-term physical and mechanical properties and microstructures of fly-ash-based geopolymer composite incorporating carbide slag. Materials 2021, 14, 6692. [Google Scholar] [CrossRef] [PubMed]
  215. Shinde, P.; Patankar, S.; Sayyad, A. Investigation on effects of fineness of flyash and alkaline ratio on mechanical properties of geopolymer concrete. Res. Eng. Struct. Mater. 2018, 4, 61–71. [Google Scholar] [CrossRef]
  216. Nikoloutsopoulos, N.; Sotiropoulou, A.; Kakali, G.; Tsivilis, S. The Effect of Solid/Liquid Ratio on Setting Time, Workability and Compressive Strength of Fly Ash Based Geopolymers. Mater. Today Proc. 2018, 5, 27441–27445. [Google Scholar] [CrossRef]
  217. Arafa, S.A.; Ali, A.Z.M.; Awal, A.S.M.A.; Loon, L.Y. Optimum mix for fly ash geopolymer binder based on workability and compressive strength. In Proceedings of the IOP Conference Series: Earth and Environmental Science, Langkawi, Malaysia, 4–5 December 2017; Institute of Physics Publishing: Bristol, UK, 2018. [Google Scholar] [CrossRef]
  218. Albitar, M.; Visintin, P.; Ali, M.S.M.; Drechsler, M. Assessing behaviour of fresh and hardened geopolymer concrete mixed with class-F fly ash. KSCE J. Civ. Eng. 2015, 19, 1445–1455. [Google Scholar] [CrossRef]
  219. Wang, W.; Fan, C.; Wang, B.; Zhang, X.; Liu, Z. Workability, rheology, and geopolymerization of fly ash geopolymer: Role of alkali content, modulus, and water–binder ratio. Constr. Build. Mater. 2023, 367, 130357. [Google Scholar] [CrossRef]
  220. Luga, E.; Atis, C.D. Optimization of heat cured fly ash/slag blend geopolymer mortars designed by “Combined Design” method: Part 1. Constr. Build. Mater. 2018, 178, 393–404. [Google Scholar] [CrossRef]
  221. Hadi, M.N.S.; Zhang, H.; Parkinson, S. Optimum mix design of geopolymer pastes and concretes cured in ambient condition based on compressive strength, setting time and workability. J. Build. Eng. 2019, 23, 301–313. [Google Scholar] [CrossRef]
  222. Ketana, N.S.; Reddy, V.S.; Rao, M.V.S.; Shrihari, S. Effect of various parameters on the workability and strength properties of geopolymer concrete. Proc. E3S Web Conf. 2021, 309, 01102. [Google Scholar] [CrossRef]
  223. Naghizadeh, A.; Ekolu, S.O. Method for comprehensive mix design of fly ash geopolymer mortars. Constr. Build. Mater. 2019, 202, 704–717. [Google Scholar] [CrossRef]
  224. Khodr, M.; Law, D.W.; Gunasekara, C.; Setunge, S.; Brkljaca, R. Compressive strength and microstructure evolution of low calcium brown coal fly ash-based geopolymer. J. Sustain. Cem. Based Mater. 2020, 9, 17–34. [Google Scholar] [CrossRef]
  225. Romadhon, E.S. The influence of low alkaline activator on the compressive strength and workability of geopolymer concrete. In Proceedings of the IOP Conference Series: Earth and Environmental Science, Rapperswil, Switzerland, 3–12 October 2022; Institute of Physics: London, UK, 2023. [Google Scholar] [CrossRef]
  226. Paul, E. Influence of superplasticizer on workability and strength of ambient cured alkali activated mortar. Clean. Mater. 2022, 6, 100152. [Google Scholar] [CrossRef]
  227. Deb, P.S.; Nath, P.; Sarker, P.K. The effects of ground granulated blast-furnace slag blending with fly ash and activator content on the workability and strength properties of geopolymer concrete cured at ambient temperature. Mater. Des. 2014, 62, 32–39. [Google Scholar] [CrossRef]
  228. Nishanth, L.; Patil, D.N.N. Experimental evaluation on workability and strength characteristics of self-consolidating geopolymer concrete based on GGBFS, flyash and alccofine. Mater. Today Proc. 2022, 59, 51–57. [Google Scholar] [CrossRef]
  229. Nagajothi, S.; Elavenil, S. Parametric studies on the workability and compressive strength properties of geopolymer concrete. J. Mech. Behav. Mater. 2018, 27, 20180019. [Google Scholar] [CrossRef]
  230. Nuruddin, M.F.; Demie, S.; Ahmed, M.F.; Shafiq, N. Effect of Superplasticizer NaOHMolarity on Workability Compressive Strength Microstructure Properties of Self-Compacting Geopolymer Concrete. Int. J. Civ. Environ. Eng. 2011, 2, 122–129. [Google Scholar]
  231. Al-Majidi, M.H.; Lampropoulos, A.; Cundy, A. Effect of Alkaline Activator, Water. Superplasticiser and Slag Contents on the Compressive Strength and Workability of Slag-Fly Ash Based Geopolymer Mortar Cured under Ambient Temperature. Int. J. Civ. Environ. Struct. Constr. Archit. Eng. 2016, 10, 308–312. [Google Scholar]
  232. Rajput, B.S.; Rajawat, S.P.S.; Jain, G. Effect of curing conditions on the compressive strength of fly ash-based geopolymer concrete. Mater. Today Proc. 2023, 103, 32–38. [Google Scholar] [CrossRef]
  233. Song, W.; Zhu, Z.; Peng, Y.; Wan, Y.; Xu, X.; Pu, S.; Song, S.; Wei, Y. Effect of steel slag on fresh, hardened and microstructural properties of high-calcium fly ash based geopolymers at standard curing condition. Constr. Build. Mater. 2019, 229, 116933. [Google Scholar] [CrossRef]
  234. Junaid, M.T.; Kayali, O.; Khennane, A.; Black, J. A mix design procedure for low calcium alkali activated fly ash-based concretes. Constr. Build. Mater. 2015, 79, 301–310. [Google Scholar] [CrossRef]
  235. Rangarao, M.L.S.; Pradhan, B. Effect of chloride and blend of chloride and sulphate salts on workability, early strength and microstructure of FA-GGBS geopolymer concrete. Mater. Today Proc. 2022, 65, 3907–3911. [Google Scholar] [CrossRef]
  236. Chithambaram, S.J.; Kumar, S.; Prasad, M.M.; Adak, D. Effect of parameters on the compressive strength of fly ash based geopolymer concrete. Struct. Concr. 2018, 19, 1202–1209. [Google Scholar] [CrossRef]
  237. Topark-Ngarm, P.; Chindaprasirt, P.; Sata, V. Setting Time, Strength, and Bond of High-Calcium Fly Ash Geopolymer Concrete. J. Mater. Civ. Eng. 2015, 27, 04014198. [Google Scholar] [CrossRef]
  238. Naenudon, S.; Wongsa, A.; Ekprasert, J.; Sata, V.; Chindaprasirt, P. Enhancing the properties of fly ash-based geopolymer concrete using recycled aggregate from waste ceramic electrical insulator. J. Build. Eng. 2023, 68, 106132. [Google Scholar] [CrossRef]
  239. Nath, P.; Sarker, P.K. Effect of GGBFS on setting, workability and early strength properties of fly ash geopolymer concrete cured in ambient condition. Constr. Build. Mater. 2014, 66, 163–171. [Google Scholar] [CrossRef]
  240. Waqas, R.M.; Butt, F.; Zhu, X.; Jiang, T.; Tufail, R.F. A comprehensive study on the factors affecting the workability and mechanical properties of ambient cured fly ash and slag based geopolymer concrete. Appl. Sci. 2021, 11, 8722. [Google Scholar] [CrossRef]
  241. Özbayrak, A.; Kucukgoncu, H.; Atas, O.; Aslanbay, H.H.; Aslanbay, Y.G.; Altun, F. Determination of stress-strain relationship based on alkali activator ratios in geopolymer concretes and development of empirical formulations. Structures 2023, 48, 2048–2061. [Google Scholar] [CrossRef]
  242. Fang, G.; Ho, W.K.; Tu, W.; Zhang, M. Workability and mechanical properties of alkali-activated fly ash-slag concrete cured at ambient temperature. Constr. Build. Mater. 2018, 172, 476–487. [Google Scholar] [CrossRef]
  243. Wardhono, A.; Law, D.W.; Sutikno; Dani, H. The effect of slag addition on strength development of Class C fly ash geopolymer concrete at normal temperature. In Proceedings of the AIP Conference Proceedings, East Java, Indonesia, 8–9 August 2017; American Institute of Physics Inc.: College Park, MD, USA, 2017. [Google Scholar] [CrossRef]
  244. Nematollahi, B.; Sanjayan, J. Effect of different superplasticizers and activator combinations on workability and strength of fly ash based geopolymer. Mater. Des. 2014, 57, 667–672. [Google Scholar] [CrossRef]
  245. Li, J.; Dang, X.; Zhang, J.; Yi, P.; Li, Y. Mechanical Properties of Fly Ash-Slag Based Geopolymer for Repair of Road Subgrade Diseases. Polymers 2023, 15, 309. [Google Scholar] [CrossRef]
  246. Sunarsih, E.S.; As’ad, S.; Sam, A.R.M.; Kristiawan, S.A. Properties of Fly Ash-Slag-Based Geopolymer Concrete with Low Molarity Sodium Hydroxide. Civ. Eng. J. 2023, 9, 381–392. [Google Scholar] [CrossRef]
  247. Assi, L.N.; Deaver, E.; Elbatanouny, M.K.; Ziehl, P. Investigation of early compressive strength of fly ash-based geopolymer concrete. Constr. Build. Mater. 2016, 112, 807–815. [Google Scholar] [CrossRef]
  248. Cho, Y.K.; Yoo, S.W.; Jung, S.H.; Lee, K.M.; Kwon, S.J. Effect of Na2O content, SiO2/Na2O molar ratio, and curing conditions on the compressive strength of FA-based geopolymer. Constr. Build. Mater. 2017, 145, 253–260. [Google Scholar] [CrossRef]
  249. Adam, A.A.; Horianto, X. The effect of temperature and duration of curing on the strength of fly ash based geopolymer mortar. Procedia Eng. 2014, 95, 410–414. [Google Scholar] [CrossRef]
  250. Ahmad, M.; Rashid, K.; Tariq, Z.; Ju, M. Utilization of a novel artificial intelligence technique (ANFIS) to predict the compressive strength of fly ash-based geopolymer. Constr. Build. Mater. 2021, 301, 124251. [Google Scholar] [CrossRef]
  251. Samantasinghar, S.; Singh, S.P. Effect of synthesis parameters on compressive strength of fly ash-slag blended geopolymer. Constr. Build. Mater. 2018, 170, 225–234. [Google Scholar] [CrossRef]
  252. Sukmak, P.; Horpibulsuk, S.; Shen, S.L. Strength development in clay-fly ash geopolymer. Constr. Build. Mater. 2013, 40, 566–574. [Google Scholar] [CrossRef]
  253. Ahmed, H.U.; Mohammed, A.S.; Qaidi, S.M.A.; Faraj, R.H.; Sor, N.H.; Mohammed, A.A. Compressive strength of geopolymer concrete composites: A systematic comprehensive review, analysis and modeling. Eur. J. Environ. Civ. Eng. 2023, 27, 1383–1428. [Google Scholar] [CrossRef]
  254. Guades, E.J. Experimental investigation of the compressive and tensile strengths of geopolymer mortar: The effect of sand/fly ash (S/FA) ratio. Constr. Build. Mater. 2016, 127, 484–493. [Google Scholar] [CrossRef]
  255. Hoy, M.; Horpibulsuk, S.; Arulrajah, A. Strength development of Recycled Asphalt Pavement—Fly ash geopolymer as a road construction material. Constr. Build. Mater. 2016, 117, 209–219. [Google Scholar] [CrossRef]
  256. Suksiripattanapong, C.; Horpibulsuk, S.; Chanprasert, P.; Sukmak, P.; Arulrajah, A. Compressive strength development in fly ash geopolymer masonry units manufactured from water treatment sludge. Constr. Build. Mater. 2015, 82, 20–30. [Google Scholar] [CrossRef]
  257. Mustafa, A.M.; Bakri, A.; Kamarudin, H.; Bnhussain, M.; Rafiza, A.R.; Zarina, Y. Effect of Na2SiO3/NaOHRatios NaOHMolarities on Compressive Strength of Fly-Ash-Based Geopolymer. ACI Mater. J. 2012, 109, 503–508. [Google Scholar]
  258. Lim, N.H.A.S.; Samadi, M.; Ariffin, N.F.; Hussin, M.W.; Bhutta, M.A.R.; Sarbini, N.N.; Khalid, N.H.A.; Aminuddin, E. Effect of Curing Conditions on Compressive Strength of FA-POFA-based Geopolymer Mortar. In Proceedings of the IOP Conference Series: Materials Science and Engineering, Istanbul, Turkey, 8–9 August 2018; Institute of Physics Publishing: Bristol, UK, 2018. [Google Scholar] [CrossRef]
  259. Pangdaeng, S.; Phoo-ngernkham, T.; Sata, V.; Chindaprasirt, P. Influence of curing conditions on properties of high calcium fly ash geopolymer containing Portland cement as additive. Mater. Des. 2014, 53, 269–274. [Google Scholar] [CrossRef]
  260. Arulrajah, A.; Kua, T.A.; Horpibulsuk, S.; Phetchuay, C.; Suksiripattanapong, C.; Du, Y.J. Strength and microstructure evaluation of recycled glass-fly ash geopolymer as low-carbon masonry units. Constr. Build. Mater. 2016, 114, 400–406. [Google Scholar] [CrossRef]
  261. Zhang, L.; Chen, Z.; Chen, R.; Zhu, S.; Lin, J.; Tai, P. Compressive strength of fly ash based geopolymer utilizing waste completely decomposed granite. Case Stud. Constr. Mater. 2023, 19, e02667. [Google Scholar] [CrossRef]
  262. Ahmed, H.U.; Mohammed, A.A.; Rafiq, S.; Mohammed, A.S.; Mosavi, A.; Sor, N.H.; Qaidi, S.M.A. Compressive strength of sustainable geopolymer concrete composites: A state-of-the-art review. Sustainability 2021, 13, 13502. [Google Scholar] [CrossRef]
  263. Hadi, M.N.S.; Al-Azzawi, M.; Yu, T. Effects of fly ash characteristics and alkaline activator components on compressive strength of fly ash-based geopolymer mortar. Constr. Build. Mater. 2018, 175, 41–54. [Google Scholar] [CrossRef]
  264. Meesala, C.R.; Verma, N.K.; Kumar, S. Critical review on fly-ash based geopolymer concrete. Struct. Concr. 2020, 21, 1013–1028. [Google Scholar] [CrossRef]
  265. Cai, J.; Li, X.; Tan, J.; Vandevyvere, B. Thermal and compressive behaviors of fly ash and metakaolin-based geopolymer. J. Build. Eng. 2020, 30, 101307. [Google Scholar] [CrossRef]
  266. Elyamany, H.E.; Elmoaty, A.E.M.A.; Elshaboury, A.M. Setting time and 7-day strength of geopolymer mortar with various binders. Constr. Build. Mater. 2018, 187, 974–983. [Google Scholar] [CrossRef]
  267. Li, Z.; Dong, H.; Zhao, X.; Wang, K.; Gao, X. Utilisation of Bayer red mud for high-performance geopolymer: Competitive roles of different activators. Case Stud. Constr. Mater. 2025, 23, e05047. [Google Scholar] [CrossRef]
  268. Puligilla, S.; Mondal, P. Role of slag in microstructural development and hardening of fly ash-slag geopolymer. Cem. Concr. Res. 2013, 43, 70–80. [Google Scholar] [CrossRef]
  269. Rai, B.; Roy, L.B.; Rajjak, M. A statistical investigation of different parameters influencing compressive strength of fly ash induced geopolymer concrete. Struct. Concr. 2018, 19, 1268–1279. [Google Scholar] [CrossRef]
  270. Hamidi, F.; Aslani, F.; Valizadeh, A. Compressive and tensile strength fracture models for heavyweight geopolymer concrete. Eng. Fract. Mech. 2020, 231, 107023. [Google Scholar] [CrossRef]
  271. Cui, Y.; Gao, K.; Zhang, P. Experimental and statistical study on mechanical characteristics of geopolymer concrete. Materials 2020, 13, 1651. [Google Scholar] [CrossRef] [PubMed]
  272. Nguy, H.H.; Lương, Q.H.; Choi, J.I.; Ranade, R.; Li, V.C.; Lee, B.Y. Ultra-ductile behavior of fly ash-based engineered geopolymer composites with a tensile strain capacity up to 13.7%. Cem. Concr. Compos. 2021, 122, 104133. [Google Scholar] [CrossRef]
  273. Abbas, S.N.; Qureshi, M.I.; Abid, M.M.; Zia, A.; Tariq, M.A.U.R. An Investigation of Mechanical Properties of Fly Ash Based Geopolymer and Glass Fibers Concrete. Sustainability 2022, 14, 10489. [Google Scholar] [CrossRef]
  274. Kan, L.L.; Wang, W.S.; Liu, W.D.; Wu, M. Development and characterization of fly ash based PVA fiber reinforced Engineered Geopolymer Composites incorporating metakaolin. Cem. Concr. Compos. 2020, 108, 103521. [Google Scholar] [CrossRef]
  275. Pratap, B.; Mondal, S.; Rao, B.H. Development of geopolymer concrete using fly ash and phosphogypsum as a pavement composite material. Mater. Today Proc. 2023, 93, 35–40. [Google Scholar] [CrossRef]
  276. Lao, J.C.; Ma, R.Y.; Xu, L.Y.; Li, Y.; Shen, Y.N.; Yao, J.; Wang, Y.S.; Xie, T.Y.; Huang, B.T. Fly ash-dominated High-Strength Engineered/Strain-Hardening Geopolymer Composites (HS-EGC/SHGC): Influence of alkalinity and environmental assessment. J. Clean. Prod. 2024, 447, 141182. [Google Scholar] [CrossRef]
  277. Ashfaq, M.H.; Sharif, M.B.; Irfan-ul-Hassan, M.; Sahar, U.U.; Akmal, U.; Mohamed, A. Up-scaling of fly ash-based geopolymer concrete to investigate the binary effect of locally available metakaolin with fly ash. Heliyon 2024, 10, e26331. [Google Scholar] [CrossRef] [PubMed]
  278. Olivia, M.; Nikraz, H. Properties of fly ash geopolymer concrete designed by Taguchi method. Mater. Des. 2012, 36, 191–198. [Google Scholar] [CrossRef]
  279. Kumar, S.; Pradeepa, J.; Ravindra, P.M. Experimental Investigations on Optimal Strength Parameters of Fly Ash Based Geopolymer Concrete. Int. J. Civ. Struct. Civ. Eng. Res. 2013, 2, 143. [Google Scholar]
  280. Rathod, G.; Sanni, P.S.H. Microstructure Studies on the Effect of Alkaline Activators to Flyash Ratio on Geopolymer Concrete. Int. J. Res. Appl. Sci. Eng. Technol. 2023, 11, 1220–1225. [Google Scholar] [CrossRef]
  281. El Abd, A.; Taman, M.; Behiry, R.N.; El-Naggar, M.R.; Eissa, M.; Hassan, A.M.A.; Bar, W.A.; Mongy, T.; Osman, M.; Hassan, A.; et al. Neutron imaging of moisture transport, water absorption characteristics and strength properties for fly ash/slag blended geopolymer mortars: Effect of drying temperature. Constr. Build. Mater. 2024, 449, 138436. [Google Scholar] [CrossRef]
  282. Sitarz, M.; Hager, I.; Choińska, M. Evolution of mechanical properties with time of fly-ash-based geopolymer mortars under the effect of granulated ground blast furnace slag addition. Energies 2020, 13, 1135. [Google Scholar] [CrossRef]
  283. Guo, X.; Li, S. Effect of Curing Age on Tensile Properties of Fly Ash Based Engineered Geopolymer Composites (FA-EGC) by Uniaxial Tensile Test and Ultrasonic Pulse Velocity Method. J. Wuhan Univ. Technol. Mater. Sci. Ed. 2023, 38, 1319–1325. [Google Scholar] [CrossRef]
  284. Bezabih, T.; Kanali, C.; Thuo, J. Effects of teff straw ash on the mechanical and microstructural properties of ambient cured fly ash-based geopolymer mortar for onsite applications. Results Eng. 2023, 18, 101123. [Google Scholar] [CrossRef]
  285. Adak, D.; Sarkar, M.; Mandal, S. Effect of nano-silica on strength and durability of fly ash based geopolymer mortar. Constr. Build. Mater. 2014, 70, 453–459. [Google Scholar] [CrossRef]
  286. Azarsa, P.; Gupta, R. Comparative study involving effect of curing regime on elastic modulus of geopolymer concrete. Buildings 2020, 10, 101. [Google Scholar] [CrossRef]
  287. Bellum, R.R.; Muniraj, K.; Indukuri, C.S.R.; Madduru, S.R.C. Investigation on Performance Enhancement of Fly ash-GGBFS Based Graphene Geopolymer Concrete. J. Build. Eng. 2020, 32, 101659. [Google Scholar] [CrossRef]
  288. Jagadesh, P.; Nagarajan, V.; Karthik Arunachalam, K. Effect of nano titanium di oxide on mechanical properties of fly ash and ground granulated blast furnace slag based geopolymer concrete. J. Build. Eng. 2022, 61, 105235. [Google Scholar] [CrossRef]
  289. Deb, P.S.; Khan, M.N.N.; Sarker, P.K.; Barbhuiya, S. Nanomechanical characterization of ambient-cured fly ash geopolymers containing nanosilica. J. Sustain. Cem. Based Mater. 2022, 11, 164–174. [Google Scholar] [CrossRef]
  290. Aslanbay, Y.G.; Aslanbay, H.H.; Özbayrak, A.; Kucukgoncu, H.; Atas, O. Comprehensive analysis of experimental and numerical results of bond strength and mechanical properties of fly ash based GPC and OPC concrete. Constr. Build. Mater. 2024, 416, 135175. [Google Scholar] [CrossRef]
  291. Bellum, R.R. Influence of steel and PP fibers on mechanical and microstructural properties of fly ash-GGBFS based geopolymer composites. Ceram. Int. 2022, 48, 6808–6818. [Google Scholar] [CrossRef]
  292. Fang, Y.; Wang, C.; Yang, H.; Chen, J.; Dong, Z.; Li, L.Y. Development of a ternary high-temperature resistant geopolymer and the deterioration mechanism of its concrete after heat exposure. Constr. Build. Mater. 2024, 449, 138291. [Google Scholar] [CrossRef]
  293. Nguyen, K.T.; Ahn, N.; Le, T.A.; Lee, K. Theoretical and experimental study on mechanical properties and flexural strength of fly ash-geopolymer concrete. Constr. Build. Mater. 2016, 106, 65–77. [Google Scholar] [CrossRef]
  294. Wang, J.; Wu, D.; Chen, K.; Mao, N.; Zhang, Z. Effect of Ca content on the synthesis and properties of FA-GGBFS geopolymer: Combining experiments and molecular dynamics simulation. J. Build. Eng. 2024, 94, 109908. [Google Scholar] [CrossRef]
  295. Zhao, R.; Yuan, Y.; Cheng, Z.; Wen, T.; Li, J.; Li, F.; Ma, Z.J. Freeze-thaw resistance of Class F fly ash-based geopolymer concrete. Constr. Build. Mater. 2019, 222, 474–483. [Google Scholar] [CrossRef]
  296. Phoo-ngernkham, T.; Chindaprasirt, P.; Sata, V.; Hanjitsuwan, S.; Hatanaka, S. The effect of adding nano-SiO2 and nano-Al2O3 on properties of high calcium fly ash geopolymer cured at ambient temperature. Mater. Des. 2014, 55, 58–65. [Google Scholar] [CrossRef]
  297. Rovnaník, P.; Šimonová, H.; Topolář, L.; Schmid, P.; Keršner, Z. Effect of Carbon Nanotubes on the Mechanical Fracture Properties of Fly Ash Geopolymer. Procedia Eng. 2016, 151, 321–328. [Google Scholar] [CrossRef]
  298. Nath, P.; Sarker, P.K. Flexural strength and elastic modulus of ambient-cured blended low-calcium fly ash geopolymer concrete. Constr. Build. Mater. 2017, 130, 22–31. [Google Scholar] [CrossRef]
  299. Bellum, R.R.; Muniraj, K.; Madduru, S.R.C. Investigation on modulus of elasticity of fly ash-ground granulated blast furnace slag blended geopolymer concrete. Mater. Today Proc. 2020, 27, 718–723. [Google Scholar] [CrossRef]
  300. Romadhon, E.S.; Antonius, S. Mechanical Properties of Geopolymer Concrete Containing Low-Alkaline Activator. Ann. Chim. Sci. Mater. 2022, 46, 273–279. [Google Scholar] [CrossRef]
  301. Morsy, A.M.; Ragheb, A.M.; Shalan, A.H.; Mohamed, O.H. Mechanical Characteristics of GGBFS/FA-Based Geopolymer Concrete and Its Environmental Impact. Pract. Period. Struct. Des. Constr. 2022, 27, 04022017. [Google Scholar] [CrossRef]
  302. Yilmazoglu, A.; Yildirim, S.T.; Yıldız, S.; Behçet, Ö.F. Effect of GGBS Content and Water/Geopolymer Solid Ratio on the Mechanical, Elevated Temperature Resistance, and Sorptivity Properties of FA/GGBS-Based Geopolymer Concrete. J. Mater. Civ. Eng. 2024, 36, 04024032. [Google Scholar] [CrossRef]
  303. Nikoloutsopoulos, N.; Sotiropoulou, A.; Kakali, G.; Tsivilis, S. Physical and Mechanical Properties of Fly Ash Based Geopolymer Concrete Compared to Conventional Concrete. Buildings 2021, 11, 178. [Google Scholar] [CrossRef]
  304. Kuri, J.C.; Sarker, P.K.; Shaikh, F.U.A. Sulphuric acid resistance of ground ferronickel slag blended fly ash geopolymer mortar. Constr. Build. Mater. 2021, 313, 125505. [Google Scholar] [CrossRef]
  305. Wang, H.T.; Yang, X.Y.; Wang, M.; Qu, C. Influence of Silicon-Aluminium Modified Materials on Durability of Fly Ash-Based Geopolymers. Mater. Sci. Forum 2023, 1103, 139–154. [Google Scholar] [CrossRef]
  306. Yang, W.; Zhu, P.; Liu, H.; Wang, X.; Ge, W.; Hua, M. Resistance to sulfuric acid corrosion of geopolymer concrete based on different binding materials and Alkali concentrations. Materials 2021, 14, 7109. [Google Scholar] [CrossRef] [PubMed]
  307. Ojha, A.; Aggarwal, P. Durability performance of low calcium Flyash-Based geopolymer concrete. Structures 2023, 54, 956–963. [Google Scholar] [CrossRef]
  308. Sharmila, S.; Natarajan, N.; Dineshkumar, S. Analysis of microstructural properties and durability characteristics of geopolymer mortar. Mater. Today Proc. 2023, 93, 217–222. [Google Scholar] [CrossRef]
  309. Saptamongkol, A.; Sata, V.; Wongsa, A.; Kroehong, W.; Ekprasert, J.; Chindaprasirt, P. Hybrid geopolymer paste from high calcium fly ash and glass wool: Mechanical, microstructure, and sulfuric acid and magnesium sulfate resistance characteristics. J. Build. Eng. 2023, 76, 107245. [Google Scholar] [CrossRef]
  310. Aiken, T.A.; Kwasny, J.; Sha, W.; Soutsos, M.N. Effect of slag content and activator dosage on the resistance of fly ash geopolymer binders to sulfuric acid attack. Cem. Concr. Res. 2018, 111, 23–40. [Google Scholar] [CrossRef]
  311. Ganeshan, M.; Venkataraman, S. Durability and microstructural studies on fly ash blended self-compacting geopolymer concrete. Eur. J. Environ. Civ. Eng. 2021, 25, 2074–2088. [Google Scholar] [CrossRef]
  312. Agnihotri, A.; Ramana, P.V. GGBS: Fly-Ash evaluation and mechanical properties within high strength concrete. Mater. Today Proc. 2021, 50, 2404–2410. [Google Scholar] [CrossRef]
  313. Ajay, A.; Ramaswamy, K.P.; Thomas, A.V. Mechanical properties of alkali-activated mortar in organic acid environment. In Proceedings of the IOP Conference Series: Materials Science and Engineering, Chennai, India, 16–17 July 2020; IOP Publishing Ltd.: Bristol, UK, 2020. [Google Scholar] [CrossRef]
  314. Lavanya, G.; Jegan, J. Durability Study on High Calcium Fly Ash Based Geopolymer Concrete. Adv. Mater. Sci. Eng. 2015, 2015, 731056. [Google Scholar] [CrossRef]
  315. Ariffin, M.A.M.; Bhutta, M.A.R.; Hussin, M.W.; Tahir, M.M.; Aziah, N. Sulfuric acid resistance of blended ash geopolymer concrete. Constr. Build. Mater. 2013, 43, 80–86. [Google Scholar] [CrossRef]
  316. Thokchom, S.; Ghosh, P.; Ghosh, S. Durability of fly ash geopolymer mortars in nitric ACID—Effect of alkali (Na2O) content. J. Civ. Eng. Manag. 2011, 17, 393–399. [Google Scholar] [CrossRef]
  317. Nnaemeka, O.F.; Singh, N.B. Durability properties of geopolymer concrete made from fly ash in presence of Kaolin. Mater. Today Proc. 2019, 29, 781–784. [Google Scholar] [CrossRef]
  318. Okoye, F.N.; Prakash, S.; Singh, N.B. Durability of fly ash based geopolymer concrete in the presence of silica fume. J. Clean. Prod. 2017, 149, 1062–1067. [Google Scholar] [CrossRef]
  319. Nguyen, K.T.; Lee, Y.H.; Lee, J.; Ahn, N. Acid resistance and curing properties for green fly ash-geopolymer concrete. J. Asian Arch. Build. Eng. 2013, 12, 317–322. [Google Scholar] [CrossRef]
  320. Xie, J.; Zhao, J.; Wang, J.; Wang, C.; Huang, P.; Fang, C. Sulfate resistance of recycled aggregate concrete with GGBS and fly ash-based geopolymer. Materials 2019, 12, 1247. [Google Scholar] [CrossRef]
  321. Patrisia, Y.; Law, D.W.; Gunasekara, C.; Wardhono, A. Fly ash geopolymer concrete durability to sulphate, acid and peat attack. Proc. MATEC Web Conf. 2022, 364, 02003. [Google Scholar] [CrossRef]
  322. Alyousef, R.; Ebid, A.A.K.; Huseien, G.F.; Mohammadhosseini, H.; Alabduljabbar, H.; Ngian, S.P.; Mohamed, A.M. Effects of Sulfate and Sulfuric Acid on Efficiency of Geopolymers as Concrete Repair Materials. Gels 2022, 8, 53. [Google Scholar] [CrossRef]
  323. Saavedra, W.G.V.; Angulo, D.E.; de Gutiérrez, R.M. Fly Ash Slag Geopolymer Concrete: Resistance to Sodium and Magnesium Sulfate Attack. J. Mater. Civ. Eng. 2016, 28, 04016148. [Google Scholar] [CrossRef]
  324. Rattanachan, S.T.; Dangwiriyakul, R.; Thongsri, O.; Thaitalay, P.; Srisuwan, S. Resistance of Chloride and Sulfate of Geopolymer Mortar Based on Fly Ash and Silica Fume as Alternative Raw Materials. Suranaree J. Sci. Technol. 2023, 30, 030147. [Google Scholar] [CrossRef]
  325. Sastry, K.V.S.G.K.; Sahitya, P.; Ravitheja, A. Influence of nano TiO2 on strength and durability properties of geopolymer concrete. Mater. Today Proc. 2021, 45, 1017–1025. [Google Scholar] [CrossRef]
  326. Ariffin, M.A.M.; Hussin, M.W. Chloride Resistance of Blended Ash Geopolymer Concrete. J. Civ. Eng. Sci. Technol. 2025, 6, 23–33. [Google Scholar] [CrossRef]
  327. Amin, N.U.; Nawab, L.; Ghani, U. Synthesis and characterization of chloride resistant cement from industrial waste through geopolymerization. J. Clean. Prod. 2017, 156, 577–580. [Google Scholar] [CrossRef]
  328. Singh, R.P.; Vanapalli, K.R.; Jadda, K.; Mohanty, B. Durability assessment of fly ash, GGBS, and silica fume based geopolymer concrete with recycled aggregates against acid and sulfate attack. J. Build. Eng. 2024, 82, 108354. [Google Scholar] [CrossRef]
  329. Mohamed, O.A.; Al-Khattab, R.; Al-Hawat, W. Resistance to acid degradation, sorptivity, and setting time of geopolymer mortars. Front. Struct. Civ. Eng. 2022, 16, 781–791. [Google Scholar] [CrossRef]
  330. Çevik, A.; Alzeebaree, R.; Humur, G.; Niş, A.; Gülşan, M.E. Effect of nano-silica on the chemical durability and mechanical performance of fly ash based geopolymer concrete. Ceram. Int. 2018, 44, 12253–12264. [Google Scholar] [CrossRef]
  331. Morla, P.; Gupta, R.; Azarsa, P.; Sharma, A. Corrosion evaluation of geopolymer concrete made with fly ash and bottom ash. Sustainability 2021, 13, 398. [Google Scholar] [CrossRef]
  332. Farhana, F.; Kamarudin, H.; Rahmat, A.; Al, M.M.; Abdullah, B. A Study on Corrosion Behavior of Reinforcement Bar Embedded in Geopolymer Paste by Open Circuit Potential. Aust. J. Basic. Appl. Sci. 2014, 7, 230–235. [Google Scholar]
  333. Reddy, D.V.; Edouard, J.-B.; Sobhan, K. Durability of Fly Ash-Based Geopolymer Structural Concrete in the Marine Environment. Mater. Civ. Eng. 2013, 25, 781–787. [Google Scholar] [CrossRef]
  334. Rajagopalan, G. Durability of alumina silicate concrete based on slag/fly ash blends against corrosion. Eng. Constr. Arch. Manag. 2019, 26, 1641–1651. [Google Scholar] [CrossRef]
  335. Kumar, S.; Ramujee, K. Assessment of chloride ion penetration of alkali activated low calcium Fly Ash based geopolymer concrete. Key Eng. Mater. 2016, 692, 129–137. [Google Scholar] [CrossRef]
  336. Albitar, M.; Ali, M.S.M.; Visintin, P.; Drechsler, M. Durability evaluation of geopolymer and conventional concretes. Constr. Build. Mater. 2017, 136, 374–385. [Google Scholar] [CrossRef]
  337. Natali, M.E.; Manzi, S.; Carabba, L.; Chiavari, C.; Bignozzi, M.C.; Abbottoni, M.; Balbo, A.; Monticelli, C. Mechanical Performances and Corrosion Resistance of Reinforced Fly Ash Geopolymer Mortars. In 13th International Ceramics Congress—Part F; Trans Tech Publications Ltd.: Bäch, Switzerland, 2014; pp. 50–55. [Google Scholar] [CrossRef]
  338. Babaee, M.; Castel, A. Chloride-induced corrosion of reinforcement in low-calcium fly ash-based geopolymer concrete. Cem. Concr. Res. 2016, 88, 96–107. [Google Scholar] [CrossRef]
  339. Venkatesan, R.P.; Pazhani, K.C. Strength and durability properties of geopolymer concrete made with Ground Granulated Blast Furnace Slag and Black Rice Husk Ash. KSCE J. Civ. Eng. 2016, 20, 2384–2391. [Google Scholar] [CrossRef]
  340. ASTM C1898; Standard Test Methods for Determining the Chemical Resistance of Concrete Products to Acid Attack 1. ASTM International: West Conshohocken, PA, USA, 2020.
  341. ASTM C1202; Test Method for Electrical Indication of Concretes Ability to Resist Chloride Ion Penetration. ASTM International: West Conshohocken, PA, USA, 2022. [CrossRef]
  342. Kumar, C.V.; Babu, P.S.C.; Anas, M.S.M.; Varma, V.N.K.; Hemalatha, C. Assessment of Strength and Durability Parameters of Geopolymer Concrete Blocks with Different Sand Replacement Levels. Key Eng. Mater. 2023, 961, 85–93. [Google Scholar] [CrossRef]
  343. Rangan, P.R.; Tumpu, M.; Mansyur; Mabui, D.S. Assessment of Fly Ash-Rice Straw Ash-Laterite Soil Based Geopolymer Mortar Durability. Civ. Eng. J. 2023, 9, 1456–1470. [Google Scholar] [CrossRef]
  344. ASTM C1012; Test Method for Length Change of Hydraulic-Cement Mortars Exposed to a Sulfate Solution. ASTM International: West Conshohocken, PA, USA, 2024. [CrossRef]
  345. Long, T.; Zhang, H.; Chen, Y.; Li, Z.; Xu, J.; Shi, X.; Wang, Q. Effect of sulphate attack on the flexural fatigue behaviour of fly ash–based geopolymer concrete. J. Strain Anal. Eng. Des. 2018, 53, 711–718. [Google Scholar] [CrossRef]
  346. Allahverdi, A.; Hashemi, H.; Mahinroosta, M. Resistance of alkali-activated slag cement against sodium sulfate. Iran. J. Mater. Sci. Eng. 2020, 17, 23–34. [Google Scholar] [CrossRef]
  347. ASTM C876; Test Method for Corrosion Potentials of Uncoated Reinforcing Steel in Concrete. ASTM International: West Conshohocken, PA, USA, 2022. [CrossRef]
  348. Dave, S.V.; Bhogayata, A.; Arora, N.K. Mix design optimization for fresh, strength and durability properties of ambient cured alkali activated composite by Taguchi method. Constr. Build. Mater. 2021, 284, 122822. [Google Scholar] [CrossRef]
  349. Chen, Z.; Yu, J.; Nong, Y.; Yang, Y.; Zhang, H.; Tang, Y. Beyond time: Enhancing corrosion resistance of geopolymer concrete and BFRP bars in seawater. Compos. Struct. 2023, 322, 117439. [Google Scholar] [CrossRef]
  350. Hashem, F.S.; Fadel, O.; Selim, F.A.; Hassan, H.S. Examining the effect of hybrid activator on the strength and durability of slag-fine metakaolin based geopolymer cement. Sustain. Chem. Pharm. 2025, 43, 101890. [Google Scholar] [CrossRef]
  351. Eryılmaz, K.; Polat, R. Sustainable concrete production: Mechanical and durability behaviour of slag-based geopolymer containing recycled geopolymer aggregate. J. Build. Eng. 2024, 96, 110512. [Google Scholar] [CrossRef]
  352. Khoshkbijari, R.K.; Farahmandfar, A.; Zehni, N.Z.; Samimi, M.F. Properties of ground granulated blast-furnace slag-based geopolymer mortars containing glass powder, feldspar, and metakaolin under different curing conditions. Constr. Build. Mater. 2024, 435, 136753. [Google Scholar] [CrossRef]
  353. Niş, A.; Altundal, M.B. Mechanical Strength Degradation of Slag and Fly Ash Based Geopolymer Specimens Exposed to Sulfuric Acid Attack. Sigma J. Eng. Nat. Sci. 2019, 37, 917–926. [Google Scholar]
  354. Gopalakrishna, B.; Dinakar, P. The evaluation of the life cycle and corrosion properties of recycled aggregate geopolymer concrete incorporating fly ash and GGBS. J. Build. Eng. 2024, 94, 109977. [Google Scholar] [CrossRef]
  355. Assi, L.N.; Carter, K.; Deaver, E.; Ziehl, P. Review of availability of source materials for geopolymer/sustainable concrete. J. Clean. Prod. 2020, 263, 121477. [Google Scholar] [CrossRef]
  356. Orozco, C.; Babel, S.; Tangtermsirikul, S.; Sugiyama, T. Comparison of environmental impacts of fly ash and slag as cement replacement materials for mass concrete and the impact of transportation. Sustain. Mater. Technol. 2024, 39, e00796. [Google Scholar] [CrossRef]
  357. Segura, I.P.; Ranjbar, N.; Damø, A.J.; Jensen, L.S.; Canut, M.; Jensen, P.A. A review: Alkali-activated cement and concrete production technologies available in the industry. Heliyon 2023, 9, e15718. [Google Scholar] [CrossRef]
  358. Shi, X.; Zhang, C.; Liang, Y.; Luo, J.; Wang, X.; Feng, Y.; Li, Y.; Wang, Q.; Abomohra, A.E.F. Life cycle assessment and impact correlation analysis of fly ash geopolymer concrete. Materials 2021, 14, 7375. [Google Scholar] [CrossRef] [PubMed]
  359. Singh, G.V.P.B.; Prasad, V.D. Evaluating the performance and environmental impact of low calcium fly ash-based geopolymer in comparison to OPC-based concrete. Environ. Sci. Pollut. Res. 2024, 31, 66892–66910. [Google Scholar] [CrossRef]
  360. Esparham, A.; Vatin, N.I.; Kharun, M.; Hematibahar, M. A Study of Modern Eco-Friendly Composite (Geopolymer) Based on Blast Furnace Slag Compared to Conventional Concrete Using the Life Cycle Assessment Approach. Infrastructures 2023, 8, 58. [Google Scholar] [CrossRef]
  361. Habert, G.; Ouellet-Plamondon, C. Recent update on the environmental impact of geopolymers. RILEM Tech. Lett. 2016, 1, 17–23. [Google Scholar] [CrossRef]
  362. Luukkonen, T.; Abdollahnejad, Z.; Yliniemi, J.; Mastali, M.; Kinnunen, P.; Illikainen, M. Alkali-activated soapstone waste—Mechanical properties, durability, and economic prospects. Sustain. Mater. Technol. 2019, 22, e00118. [Google Scholar] [CrossRef]
  363. Heath, A.; Paine, K.; McManus, M. Minimising the global warming potential of clay based geopolymers. J. Clean. Prod. 2014, 78, 75–83. [Google Scholar] [CrossRef]
  364. Salas, D.A.; Ramirez, A.D.; Ulloa, N.; Baykara, H.; Boero, A.J. Life cycle assessment of geopolymer concrete. Constr. Build. Mater. 2018, 190, 170–177. [Google Scholar] [CrossRef]
Figure 1. Graphical representation of the amount of cement production in relation to the CO2 emission per year in billion metric tons (BMT) [9,10,11].
Figure 1. Graphical representation of the amount of cement production in relation to the CO2 emission per year in billion metric tons (BMT) [9,10,11].
Sustainability 17 11167 g001
Figure 2. PRISMA flow diagram for the systematic review.
Figure 2. PRISMA flow diagram for the systematic review.
Sustainability 17 11167 g002
Figure 3. GPC terminology illustrates the main classes of polymeric chains. Image describes the poly(sialates), which are silicon-oxo-aluminate frameworks formed by the combination of silicate and aluminate tetrahedra, cross-linked through shared oxygen atoms. The poly(sialates) serve as the fundamental building blocks of geopolymers.
Figure 3. GPC terminology illustrates the main classes of polymeric chains. Image describes the poly(sialates), which are silicon-oxo-aluminate frameworks formed by the combination of silicate and aluminate tetrahedra, cross-linked through shared oxygen atoms. The poly(sialates) serve as the fundamental building blocks of geopolymers.
Sustainability 17 11167 g003
Figure 4. Mechanism of geopolymerization. The figure describes alkalination; dissolution—depolymerization of silicate; gel formation of oligo-sialates; polycondensation; reticulation and networking; and GPC solidification, with each stage strongly dependent on the composition of the constituent precursor materials [28].
Figure 4. Mechanism of geopolymerization. The figure describes alkalination; dissolution—depolymerization of silicate; gel formation of oligo-sialates; polycondensation; reticulation and networking; and GPC solidification, with each stage strongly dependent on the composition of the constituent precursor materials [28].
Sustainability 17 11167 g004
Figure 5. Average percentage of oxide composition of various FA-based GPC with standard error bars. The mean and the standard error were computed from 25 reference data points, and the data are available in Appendix A.
Figure 5. Average percentage of oxide composition of various FA-based GPC with standard error bars. The mean and the standard error were computed from 25 reference data points, and the data are available in Appendix A.
Sustainability 17 11167 g005
Figure 6. XRD spectra showing FA-based, FA-slag-based, and slag-based GPC matrices at 28 days (where A—ZnO, B—Vaterite, C—Calcite, D—Mullite, and E—Quartz) [104]. This figure shows the mineralogy of the GPC by indicating the aluminosilicate-rich composition in FA-based and the calcium-rich composition in slag-based systems.
Figure 6. XRD spectra showing FA-based, FA-slag-based, and slag-based GPC matrices at 28 days (where A—ZnO, B—Vaterite, C—Calcite, D—Mullite, and E—Quartz) [104]. This figure shows the mineralogy of the GPC by indicating the aluminosilicate-rich composition in FA-based and the calcium-rich composition in slag-based systems.
Sustainability 17 11167 g006
Figure 7. FTIR spectra showing FA-based (FG), slag-FA-based (SFG), and slag-based (SG) GPCs [122]. The figure shows the presence of water molecules in the samples and their susceptibility to carbonation.
Figure 7. FTIR spectra showing FA-based (FG), slag-FA-based (SFG), and slag-based (SG) GPCs [122]. The figure shows the presence of water molecules in the samples and their susceptibility to carbonation.
Sustainability 17 11167 g007
Figure 8. SEM micrographs of (a) 100% FA-based GPC showing unreacted and semi-reacted FA particles surrounded by N-A-S-H gel, and (b) 60% FA–40% slag-based GPC showing a denser C-A-S-H gel matrix with fewer unreacted particles [147]. The figure shows the distinction in particles present in both FA-based and slag-based GPCs, as well as the presence of more obvious cracks in slag-based GPC due to the rate of geopolymerization.
Figure 8. SEM micrographs of (a) 100% FA-based GPC showing unreacted and semi-reacted FA particles surrounded by N-A-S-H gel, and (b) 60% FA–40% slag-based GPC showing a denser C-A-S-H gel matrix with fewer unreacted particles [147]. The figure shows the distinction in particles present in both FA-based and slag-based GPCs, as well as the presence of more obvious cracks in slag-based GPC due to the rate of geopolymerization.
Sustainability 17 11167 g008
Figure 9. TEM of FA-based GPC [75]. The figure shows phase compositions in FA-based GPC.
Figure 9. TEM of FA-based GPC [75]. The figure shows phase compositions in FA-based GPC.
Sustainability 17 11167 g009
Figure 10. TGA thermogram of (a) FA-based [171] and (b) slag-based geopolymers [172]. The figure reveals the weight loss in FA-based and slag-based geopolymers, with slag-based geopolymers showing greater overall weight loss.
Figure 10. TGA thermogram of (a) FA-based [171] and (b) slag-based geopolymers [172]. The figure reveals the weight loss in FA-based and slag-based geopolymers, with slag-based geopolymers showing greater overall weight loss.
Sustainability 17 11167 g010
Figure 11. Factors influencing physical properties of FA-based GPC with impact percentages across studies. This shows that SS/SH was identified to be the most influential factor regarding the physical properties of FA-based GPC from the studies considered.
Figure 11. Factors influencing physical properties of FA-based GPC with impact percentages across studies. This shows that SS/SH was identified to be the most influential factor regarding the physical properties of FA-based GPC from the studies considered.
Sustainability 17 11167 g011
Figure 12. Factors influencing the workability of FA-based GPC based on study frequency and reported impact. SH molarity, SS/SH ratio, and A/B ratio are the most frequently investigated parameters. However, among studies comparing multiple factors, the A/B ratio was most often identified as the most influential on workability.
Figure 12. Factors influencing the workability of FA-based GPC based on study frequency and reported impact. SH molarity, SS/SH ratio, and A/B ratio are the most frequently investigated parameters. However, among studies comparing multiple factors, the A/B ratio was most often identified as the most influential on workability.
Sustainability 17 11167 g012
Figure 13. Factors influencing compressive strength of FA-based GPC based on study frequency and reported impact. While SH molarity is the most frequently investigated parameter, the A/B ratio is most often identified as the dominant factor affecting compressive strength.
Figure 13. Factors influencing compressive strength of FA-based GPC based on study frequency and reported impact. While SH molarity is the most frequently investigated parameter, the A/B ratio is most often identified as the dominant factor affecting compressive strength.
Sustainability 17 11167 g013
Figure 14. Performance matrix of GPC and OPC concrete against aggressive agents. The figure com-pares OPC concrete with slag-based and FA–based GPC in terms of acid, chloride, sulfate, and corrosion resistance. Performance levels range from low (OPC) to excellent/superior (GPC), highlighting the superior durability of geopolymer concretes under aggressive environmental conditions.
Figure 14. Performance matrix of GPC and OPC concrete against aggressive agents. The figure com-pares OPC concrete with slag-based and FA–based GPC in terms of acid, chloride, sulfate, and corrosion resistance. Performance levels range from low (OPC) to excellent/superior (GPC), highlighting the superior durability of geopolymer concretes under aggressive environmental conditions.
Sustainability 17 11167 g014
Figure 15. (a) Relationship between resistance to acid and sulfate attack in geopolymer concrete. (b) Relationship between resistance to acid/sulfate and chloride attack, synthesized from experimental datasets in the literature. There is good correlation, indicating that improved resistance to one can significantly enhance the others.
Figure 15. (a) Relationship between resistance to acid and sulfate attack in geopolymer concrete. (b) Relationship between resistance to acid/sulfate and chloride attack, synthesized from experimental datasets in the literature. There is good correlation, indicating that improved resistance to one can significantly enhance the others.
Sustainability 17 11167 g015
Figure 16. Cost comparison for GPC and OPC mixtures for different regions. Costs are broken down by constituent materials, highlighting regional variations and the higher cost of FA–based GPC.
Figure 16. Cost comparison for GPC and OPC mixtures for different regions. Costs are broken down by constituent materials, highlighting regional variations and the higher cost of FA–based GPC.
Sustainability 17 11167 g016
Table 1. Comparison of durability properties among emerging GPC variants and OPC concrete with recommended optimization solutions.
Table 1. Comparison of durability properties among emerging GPC variants and OPC concrete with recommended optimization solutions.
Durability PropertiesRemarks
Acid ResistanceChloride ResistanceSulfate ResistanceCorrosion Resistance
Concrete TypeOPC ConcreteLowLowLowLowOPC concrete’s high calcium content reacts with acids and sulfates to form expansive compounds (gypsum and ettringite), leading to deterioration [315,317,328,336].
OPC concrete’s high porosity and permeability allow chloride ions to penetrate easily, leading to accelerated corrosion of steel, followed by spalling and structural degradation [326,332,333].
Slag-Based GPCModerateExcellentModerateGoodCalcium phases present in slag react with both acid and sulfate, generating expansive compounds, leading to deterioration of the structure [348,349,350,351].
Due to the formation of C-A-S-H phases, slag-based GPC has a denser microstructure and lower permeability, resulting in excellent resistance to chloride penetration and corrosion of steel [339,348,349,352].
FA-Based GPCHighGoodHighSuperiorDue to fly ash’s aluminosilicate framework and low calcium content, the formation of soluble or expansive compounds under acid and sulfate attack is low [317,321,328,330].
Slight porosity in FA concrete can increase permeability, making it susceptible to chloride ingress and potential corrosion initiation of steel over time [307,311,353,354].
Recommended solution to durable GPC mixIncrease silica content by adding nano-silica or graphene for stronger N-A-S-H gel.Use high SS content to improve chloride-binding capacity. Include silica-rich precursors like rice husk ash, silica fume.Use aluminosilicate-rich precursors (high-Al2O3 FA, metakaolin).Use alkaline activators with lower alkali content (lower SH molarity).
—Add nano-silica or metakaolin to refine pore structure.
Table 2. Statistics of FA, slag, FA-GP, and slag-GP (USA) [355].
Table 2. Statistics of FA, slag, FA-GP, and slag-GP (USA) [355].
Annual Concrete Production (Mt/Year)89
FA-Based GPC
FA requirement for 100% replacement of OPC (470 kg/m3)—(Mt/year)42
Annual FA available for Use—Mt/year156
Surplus (+)/Deficit (−)—Mt/year+114
FA depletion time with 100% OPC replacement (years)3214
FA Price/Ton (USD)34
FA-based GPC cement Price/Ton (USD) [357]100
FA-based GPC (Price/Ton) (USD) [357]110–120
Slag-Based GPC
Slag requirement for 100% replacement of OPC (450 kg/m3)—(Mt/year)40
Annual slag production—Mt/year20
Annual slag production available for use (Mt/year)0.98
Surplus (+)/deficit (−)—Mt/year−39
Possible % replacement of OPC with slag-GPC at current supply2%
Slag Price/Ton (USD)16–72
Slag-based GPC cement price/Ton (USD) [357]
Slag-based GPC (price/Ton) (USD) [357]130–140
% OPC Replacement Possible with current Na2SiO3 global market11%
Current Market Price of Na2SiO3/ton (USD)180–290
% OPC replacement possible with current NaOH global market75%
Current market price of NaOH/ton (USD)770–920
Cost of 1 ton OPC Concrete (USD) [357]60–65
Table 3. Environmental Impact Indicators for FA-GPC, slag-GPC, and OPC concrete (functional unit 1 m3).
Table 3. Environmental Impact Indicators for FA-GPC, slag-GPC, and OPC concrete (functional unit 1 m3).
Cradle to GateSample TypesFreshwater Toxicity (Kg 1,4-DB eq)Energy Requirement (MJ/m3)Climate Change (Kg.CO2-Eq)Human Toxicity (kg 1,4 DB-Eq)Ozone Depletion (Kg CFC-11-Eq)Acidification (kg SO2 eq)References
OPC concrete17.76355011623920.0000394[358,359]
FA-based GPC12.41–15.102400–2660646–774285–3440.00002552.57–3.07[358,359]
Change (FA GPC–OPC)(−2.66)–(−5.35)(−890)–(−1550)(−388)–(516)(−48)–(−107)−0.0000125(−0.93)–(−1.43)
OPC CML Method139.534662.7742.881872.33 × 10−131.65[360]
Slag-based GPC CML Method1001780.29172.37150.264.54 × 10−130.93[360]
Change (GPC–OPC)−39.53−2882.4−570.51−36.74+2.21 × 10−13−0.72
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mohammed, T.O.; Ul Haq, A.; Harun, M.Z.B.; Fanijo, E.O. Recent Advances in Fly Ash- and Slag-Based Geopolymer Cements. Sustainability 2025, 17, 11167. https://doi.org/10.3390/su172411167

AMA Style

Mohammed TO, Ul Haq A, Harun MZB, Fanijo EO. Recent Advances in Fly Ash- and Slag-Based Geopolymer Cements. Sustainability. 2025; 17(24):11167. https://doi.org/10.3390/su172411167

Chicago/Turabian Style

Mohammed, Taofiq O., Aman Ul Haq, Mohammad Zunaied Bin Harun, and Ebenezer O. Fanijo. 2025. "Recent Advances in Fly Ash- and Slag-Based Geopolymer Cements" Sustainability 17, no. 24: 11167. https://doi.org/10.3390/su172411167

APA Style

Mohammed, T. O., Ul Haq, A., Harun, M. Z. B., & Fanijo, E. O. (2025). Recent Advances in Fly Ash- and Slag-Based Geopolymer Cements. Sustainability, 17(24), 11167. https://doi.org/10.3390/su172411167

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop