Next Article in Journal
Increasing the Efficiency of Marine Engine Parametric Diagnostics Based on Analyses of Indicator Diagrams and Heat-Release Characteristics
Next Article in Special Issue
Large-Scale Ex Situ Tests for CO2 Storage in Coal Beds
Previous Article in Journal
Analysing Effective and Ineffective Impacts of Maintenance Strategies on Electric Power Plants: A Comprehensive Approach
Previous Article in Special Issue
A Novel Multi-Phase Strategy for Optimizing CO2 Utilization and Storage in an Oil Reservoir
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on CO2 Sequestration via Mineralization of Coal Fly Ash

1
North China Electric Power Research Institute Co., Ltd., Beijing 100045, China
2
School of Energy and Mechanical Engineering, Nanjing Normal University, Nanjing 210042, China
*
Author to whom correspondence should be addressed.
Submission received: 11 July 2023 / Revised: 24 July 2023 / Accepted: 28 July 2023 / Published: 28 August 2023
(This article belongs to the Special Issue Carbon Dioxide Capture, Utilization and Storage (CCUS) Ⅱ)

Abstract

:
Coal fly ashes (COFA) are readily available and reactive materials suitable for CO2 sequestration due to their substantial alkali components. Therefore, the onsite collaborative technology of COFA disposal and CO2 sequestration in coal-fired power plants appears to have potential. This work provides an overview of the state-of-the-art research studies in the literature on CO2 sequestration via the mineralization of COFA. The various CO2 sequestration routes of COFA are summarized, mainly including direct and indirect wet carbonation, the synthesis of porous CO2 adsorbents derived from COFA, and the development of COFA-derived inert supports for gas-solid adsorbents. The direct and indirect wet carbonation of COFA is the most concerned research technology route, which can obtain valued Ca-based by-products while achieving CO2 sequestration. Moreover, the Al and Si components rich in fly ash can be adapted to produce zeolite, hierarchical porous nano-silica, and nano-silicon/aluminum aerogels for producing highly efficient CO2 adsorbents. The prospects of CO2 sequestration technologies using COFA are also discussed. The objective of this work is to help researchers from academia and industry keep abreast of the latest progress in the study of CO2 sequestration by COFA.

1. Introduction

COFA is a type of solid waste from power plants derived from the combustion of coal [1]. Every year, a large amount of COFA is generated as the demand for cheap electricity increases [2]. As the world’s largest consumer of coal, nearly 80% of coal is consumed to produce electricity each year in China [3], producing 827 million tons of COFA in 2022. Although partial COFA is used as supplementary feedstocks to make roadbed and water-permeable bricks or as the raw materials to synthesize zeolite [4], a good deal of COFA is still accumulated or disposed of in landfills, causing harmful impacts on the local environment [5]. In addition, because heat power plants are also major sources of CO2 emissions, it is urgent to achieve CO2 reduction in coal-fired power plants via CO2 capture and storage technologies [6,7]. COFAes are inexpensive, convenient, and reactive materials suitable for mineral carbonation due to their substantial alkali and alkali earth metals [1,8]. Therefore, the onsite collaborative treatment of COFA disposal and CO2 sequestration in heat power plants can gain significant advantages [2,9]. It is similar to CO2 storage by enhanced oil recovery (EOR) and enhanced gas recovery (EGR) technologies [10,11,12], which provide additional benefits while achieving CO2 sequestration.
Generally, mineral CO2 sequestration using COFA is mainly classified into two types, including direct and indirect carbonation. Typically, direct carbonation of COFA can be accomplished by aqueous mineral carbonation or gas-solid reactions. Indirect carbonation contains the extraction of Ca2+ and Mg2+ from COFA by leaching agents (i.e., acid, ammonium salt, etc.) followed by carbonation reactions [13]. While sequestering CO2, the final carbonated products of COFA via direct carbonation can be used as construction fillers because their physical structure and leaching resistance are enhanced as well. The indirect carbonation of COFA allows for the production of pure carbonates due to the removal of impurities (i.e., silica and iron) prior to carbonate precipitation. Hence, these can enhance the economy of the carbonation treatment while decreasing the overall environmental influence of COFA. Moreover, COFA can also be used to prepare zeolites [4,14,15] and hierarchical porous nano silica [16] for CO2 adsorption. The COFA-derived inert supports for K2CO3-based, CaO-based, or amine-based sorbents for CO2 capture are also reported [17,18,19].
In this paper, the literature on CO2 sequestration using COFA has been comprehensively reviewed. The reviewed CO2 sequestration technologies majorly include direct and indirect wet carbonation, synthesis of porous CO2 adsorbents derived from COFA, and the development of COFA-derived inert supports for gas-solid adsorbents. The detailed technical route and important findings regarding the above CO2 sequestration techniques are described. Moreover, the carbonation efficiency of COFA using different sequestration technologies and the practicality of these potential technical routes are comparatively discussed. The major aim of this review is to timely grasp the development status and technical challenges of CO2 sequestration technologies using COFA and provide suggestions on the future development of these technologies.

2. CO2 Mineralization

The specific flow chart of direct and indirect CO2 mineralization using COFA is depicted in Figure 1. The CO2 sequestration is mainly accomplished by mineralization in a slurry (direct carbonation) or aqueous solution (indirect carbonation). The relevant research on direct and indirect CO2 mineralization using COFA is described in detail below.

2.1. Direct Carbonation

Direct CO2 mineralization is achieved by the reaction of COFA and CO2 either in a gaseous or aqueous phase environment. The direct CO2 mineralization route is simple and potential, and it possesses the advantages of eliminating the extraction step of reactive components and the minimal consumption of chemical reagents. As listed in Table 1, research on the direct CO2 mineralization of COFA is mainly centered on the investigation of CO2 sequestration parameters such as COFA component, carbonation temperature, carbonation pressure, carbonation time, liquid-to-solid ratio, etc. [1,9]. Moreover, there are also studies on adding calcium and magnesium ion extractants or using supercritical CO2 to promote fly ash direct mineralization to fix CO2.
As illustrated in Figure 2, the carbonation process of COFA mineralizing CO2 mainly includes three stages [20]: (I) dissolution of CO2 in fly ash slurry and ionization of H2CO3 produces CO32−; (II) Ca2+ and Mg2+ leaching into solution; (III) generation of CaCO3/MgCO3 grains that are wrapped on the surface of fly ash particles or dispersed in slurry. In fact, the carbonation reaction gradually penetrates from the outside surface to the inner core of COFA particles. Gaseous CO2 is diffused and dissolved in the liquid film surrounding the fly ash particles and reacts by leaching out Ca2+ and Mg2+. The generated carbonate products deposit on the surface of fly ash particles, leading to an uncarbonated inner core surrounded by a growing edge of carbonate products [21]. Therefore, the parameters of Ca2+/Mg2+ leaching and transportation, CO2 diffusion, carbonate product coating, and pore blockage will affect the overall rate and degree of the carbonation reaction. Among them, Ca2+/Mg2+ leaching is the rate-limiting step of COFA mineralizing CO2.
The in-situ naturally weathered COFA in a wet-dumped ash dam can achieve CO2 sequestration via carbonation to some extent; however, the carbonation reaction rate is too slow. Muriithi et al. [22] found that a period of 20 years was required to capture 6.8 wt % CO2 by in-situ natural carbonation of wet disposed ash. In contrast, the fresh COFA fixed 6.5 wt% CO2 via ex-situ accelerated carbonation (i.e., 4 Mpa, 90 °C, and liquid/solid ratio of 1 mL/g) in merely 2 h. Therefore, ex-situ accelerated carbonation of COFAes is a promising large-scale CO2 sequestration technology.
There are great discrepancies in the components of COFA due to the different coal types and the ways of flue gas desulfurization. The content of alkaline earth metals within COFAes closely affects their CO2 sequestration capacity (Cn), especially the Ca content. Ji et al. [23] dedicated themselves to studying the COFA properties’ influence on the carbonation reactions, and five types of Chinese and Australian coal combustion fly ashes were used as the samples for the study. It was found that the rich reactive Ca/Mg-bearing crystalline phases in COFA, such as CaO, Ca(OH)2, MgO, and Ca2Fe2O5 and Mg(OH)2, contributed to the increased Cn. Particularly, the Ca-bearing phases were superior to the Mg-bearing phases in improving the kinetics of carbonation reactions due to their higher reactivity with CO2. The COFA containing 32.4 wt% of CaO and 29.3 wt% of MgO exhibited the highest Cn of 132 g-CO2/kg-fly ash via accelerated carbonation (carried out at 220 °C, initial CO2 pressure of 2 MPa, 2 h and a liquid/solid ratio of 5 mL/g).
Moreover, Yuan et al. [24,25] selected four types of fly ashes derived from the subcritical, supercritical, and ultra-supercritical units, respectively, and the circulating fluidized bed, drum, and once-through boiler units. It was found that the raw fly ashes exhibited extremely low Cn of 1.8 g CO2/kg fly ash via dry mineralization. Mechanical ball milling modification could effectively enhance the Cn of fly ashes due to producing more fresh surfaces and pores within fly ashes. The CO2 sequestration capacities of wet and dry milling-modified ashes were, respectively, increased to 12.9 and 37.1 g CO2/kg fly ash [24]. In addition, wet mineralization is markedly superior to dry mineralization in improving the Cn of fly ashes. Even the raw fly ashes displayed outstanding Cn under the wet mineralization route, which was significantly better than that of mechanically modified fly ashes under dry mineralization conditions. It was mainly ascribed to the presence of water effectively accelerating the leaching of Ca2+/Mg2+ from the solid COFA particles in wet mineralization conditions [26], consequently the superior Cn. Although the water availability promotes the leaching of Ca2+/Mg2+, excess water slowed the CO2 diffusion rate due to the formed mass transfer barrier blocking the pores and cavities on the fly ash particles’ surfaces, instead reducing the carbonation efficiency. Moreover, the high reaction temperature contributed to the accelerated carbonation reaction due to the improved mass transfer rate, thermal movement of molecules, and average reaction kinetics [27,28]. However, the low CO2 solubility in water and the exothermic carbonation reaction were not conducive to CO2 sequestration at high temperatures.
To further break through the bottleneck of slow carbonation rate in the CO2 mineralization process of COFA, supercritical CO2 was adopted to enhance carbonation [24,25]. The COFA exhibited a markedly higher Cn under supercritical CO2 (54.9 g CO2/kg fly ash) compared to non-supercritical CO2 (42.3 g CO2/kg fly ash) due to the supercritical CO2 possessing the advantages of strong diffusion character and desirable permeability. The CO2 mineralization experiments of block CaO were designed by Yuan et al. [25] to measure the diffusion characteristics of supercritical and non-supercritical CO2. As illustrated in Figure 3, the supercritical CO2 (128.6 μm) exhibited a significantly deeper diffusion depth than the non-supercritical CO2 (105.5 μm) within block CaO, intuitively demonstrating the strong diffusion characteristics.
Siriwardena et al. [29] conducted experiments to measure the carbonation coefficient of cylindrical specimens (19 × 38 mm) of fly ashes. The fly ash specimens were placed in the carbonation chamber under certain conditions (reactor temperature of 40 ± 1 °C, relative humidity of 65 ± 5%, and CO2 concentration of 5 ± 0.5%). The carbonated specimens were split lengthwise, and the newly cut section was sprayed with 1% phenolphthalein indicator. The depth of carbonation could be evaluated by measuring the depth to the color boundary line between the carbonated (pH < 9.2, colorless) and uncarbonated (pH > 9.2, purple) regions, as shown in Figure 4. The following diffusion equation can describe the relationship between carbonation depth and exposure period.
X = C t + a
where X refers to the depth of carbonation (cm); C is the coefficient of carbonation ( c m d a y s ); t represents the duration of accelerated carbonation (days); and a is the empirical constant (cm). Although the COFA exhibited relatively low Cn (~41.6 g CO2/kg fly ash), the progression of carbonation in COFA was rapid, exhibiting a significantly high carbonation coefficient of 1.66 c m d a y s . It was mainly attributed to the less dense matrix in fly ash specimens, which allowed rapid diffusion of CO2 because of the generation of fewer reaction products.
The other researchers also conducted many experimental studies to reveal the effect of CO2 mineralization parameters on the Cn of coal fly ashes. However, the reported values of Cn varied in different studies due to the different selected fly ashes and CO2 mineralization parameters (Table 1) [30,31,32,33,34,35,36,37,38,39,40,41]. As shown in Figure 5, Pan et al. [32] proposed an integrated route for CO2 sequestration, fly ash stabilization, and by-product application using a high-gravity carbonation treatment method. The influence of different operating parameters on the carbonation conversion of COFA was investigated by response surface methodology. The maximal carbonation conversion of fly ash was 77.2% (i.e., Cn of 249.4 g CO2/kg fly ash) using a rotation rate of 743 rpm and a liquid/solid ratio of 18.9 mL/g at 57.3 °C. Moreover, the physicochemical properties (i.e., heavy metal leaching and volume expansion) of carbonated fly ash were upgraded, making them suitable for green materials in construction engineering.
The leaching of Ca2+ and Mg2+ into solution is an important parameter that affects carbonation efficiency. Therefore, the addition of Ca2+ and Mg2+ extractants is an effective approach to promote the leaching of Ca2+ and Mg2+ into the solution, resulting in improved Cn. Ji et al. [42] studied the effect of various 0.5 mol/L additives (i.e., Na2CO3, Na2CO3, and NaCl, NaHCO3) on the CO2 sequestration efficiency of coal fly ashes. The carbonation efficiency for different additives was ranked in the following order: 0.5 mol/L Na2CO3 > 0.5 mol/L Na2CO3 and 0.5 mol/L NaCl > 0.5 mol/L NaHCO3. The Cn was as high as 102 g CO2/kg fly ash under the conditions of 0.5 mol/L Na2CO3 and liquid/solid ratio of 10 mL/g at 275 °C and 2 MPa for 2 h. It was mainly attributed to Na2CO3, which can provide abundant CO32− in the liquid phase, which results in superior Cn. The Cn of COFA with the addition of 1 mol/L NH4Cl or sea salt was comparatively studied by Jo et al. [43]. It was suggested that the additives increase the solubility of Ca-bearing minerals, resulting in more CO2 sequestration to some extent. Therefore, the Cn in the presence of 1 mol/L NH4Cl with a liquid/solid ratio of 3 mL/g at 25 °C for 18 h was approximately 23 g CO2/kg fly ash, compared to 19 g CO2/kg fly ash in its absence.
The amine-looping process using coal fly ash could achieve efficient CO2 mineralization as well, which was confirmed in the study by Ji et al. [44,45]. They mainly investigated the CO2 fixation performance of COFA in various typical amine solutions (i.e., monoethanolamine, diethanolamine, triethanolamine, 2-amino-2-methy-1-propanol, and piperazine). It revealed that amines contributed to enhancing the mass transfer of CO2, promoting Ca2+ leaching, and generating small CaCO3 precipitation particles, especially the fly ash in 0.5 mol/L piperazine solution, which exhibited the desirable Cn of 102.9 g CO2/kg fly ash. The mechanism of amine-looping for high-efficiency CO2 carbonation in the water system using COFA was illustrated in Figure 6. The CO2 sequestration was a three-step process: mass transfer of the gaseous CO2 phase to the aqueous CO2 phase, leaching of Ca2+, and precipitation of CaCO3. Amines promoted the mass transfer of the gaseous CO2 phase to the aqueous CO2 phase by providing extra CO2 reaction pathways. Meanwhile, the protonated amines formed from carbamate generation enhanced the leaching of Ca2+, therefore being beneficial for the precipitation of CaCO3. Piperazine (a typical diamine) could offer two amino groups used to adsorb CO2, which can maintain a markedly higher CO2 concentration in the solution compared to the other four monoamines, resulting in superior CO2 sequestration performance.
To overcome the issue of the large energy consumption of absorbent regeneration for the amine-looping process, an integrated process combining CO2 absorption with mineralization was proposed [45]. A chemical method was used to regenerate the amine absorbent in the integrated process combining CO2 absorption with mineralization rather than the traditional thermal method. As shown in Figure 7, the CO2-rich solution was sent to the carbonation sink, where carbonation reactions occurred between the CaO-rich wastes and the CO2-rich solution. Afterward, the CO2 was sequestered as a solid CaCO3 phase, and the amine solvent was regenerated and returned to the absorption reactor for continuous CO2 absorption. The practical application of using COFA as a feedstock for the regeneration of absorbent in an integrated CO2 absorption-mineralization process was feasible. Results indicated that piperazine showed a larger cyclic loading of 0.42 mol/mol in the integrated process, which was approximately more than 1.1 times that of the traditional regeneration process by a thermal method.
Table 1. Summary of direct CO2 mineralization using COFA.
Table 1. Summary of direct CO2 mineralization using COFA.
ReferenceCarbonation ConditionCaO and MgO Content in Fly Ash Cn (g-CO2/kg-Fly Ash)
Shao et al. [20]liquid/solid ratio of 6 mL/g, 30 °C, initial CO2 pressure of 2 MPa, 15 min8.8 wt% Ca, 0.36 wt% Mg54.9
Muriithi et al. [22]brine, liquid/solid ratio of 1 mL/g, 90 °C, initial CO2 pressure of 4 MPa, 30 vol % CO2, 2 h9.2 wt% Ca, 2.44 wt% Mg65
in-situ natural carbonation, 20 years68
Yuan et al. [24]100 rpm, liquid/solid ratio of 100 mL/g, 80 °C, initial CO2 pressure of 1 MPa, 5 h25.83 wt% CaO, 2.17 wt% MgO42.3
100 rpm, liquid/solid ratio of 30 mL/g, 40 °C, initial CO2 pressure of 8 MPa, 5 h54.9
100 rpm, dry mineralization, 40 °C, initial CO2 pressure of 8 MPa, 5 h1.8 (raw ashes)
12.9 (dry milled ashes)
37.1 (wet milled ashes)
Yuan et al. [25]100 rpm, dry mineralization, 40 °C, initial CO2 pressure of 8 MPa, 1 h10.37 wt% CaO, 2.17 wt% MgO1.8
100 rpm, liquid/solid ratio of 10 mL/g, 40 °C, initial CO2 pressure of 3, 5 and 8 MPa, 1 h10.37 wt% CaO, 2.17 wt% MgO55 (3 MPa), 58.2 (5 MPa), 89.3 (8 MPa)
100 rpm, liquid/solid ratio of 10 mL/g, 40 °C, initial CO2 pressure of 3, 5 and 8 MPa, 1 h6.67 wt% CaO, 0.84 wt% MgO19.7 (3 MPa), 23.8 (5 MPa), 38.3 (8 MPa)
100 rpm, liquid/solid ratio of 10 mL/g, 40 °C, initial CO2 pressure of 3, 5 and 8 MPa, 1 h6.55 wt% CaO, 1.14 wt% MgO17.5 (3 MPa), 19.4 (5 MPa), 35.5 (8 MPa)
100 rpm, liquid/solid ratio of 10 mL/g, 40 °C, initial CO2 pressure of 3, 5 and 8 MPa, 1 h4.68 wt% CaO, 0.24 wt% MgO10.7 (3 MPa), 11.7 (5 MPa), 13.9 (8 MPa)
Ji et al. [23]liquid/solid ratio of 5 mL/g, 220 °C, initial CO2 pressure of 2 MPa, 2 h16.4 wt% CaO, 1.2 wt% MgO58
liquid/solid ratio of 5 mL/g, 220 °C, initial CO2 pressure of 2 MPa, 2 h9.4 wt% CaO, 27.9 wt% MgO125
liquid/solid ratio of 5 mL/g, 140 °C, initial CO2 pressure of 2 MPa, 2 h3.6 wt% CaO, 7.1 wt% MgO13
liquid/solid ratio of 5 mL/g, 140 °C, initial CO2 pressure of 2 MPa, 2 h13.4 wt% CaO, 0.5 wt% MgO26.5
liquid/solid ratio of 5 mL/g, 220 °C, initial CO2 pressure of 2 MPa, 2 h32.4 wt% CaO, 29.3 wt% MgO132
Siriwardena et al. [29]liquid/solid ratio of 0.3 mL/g, 40 °C, 10 vol % CO2, humidity of 65%, 28 days22.75 wt% CaO, 4.48 wt% MgO41.6
Back et al. [30]600 rpm, liquid/solid ratio of 20 mL/g, 75 °C, initial CO2 pressure of 0.01 MPa, 4.5 h37.3 wt% CaO,15.4 wt% MgO230
La Plante et al. [31]liquid/solid ratio of 100 mL/g, atmospheric pressure, 60 °C, 100 vol % CO2, 72 h28.5 wt% CaO, 6.6 wt% MgO95.0
Pan et al. [32]743 rpm, liquid/solid ratio of 18.9 mL/g, 57.3 °C, 15 vol % CO262.8 wt% CaO, 0.83 wt% MgO249.4
Revathy [33]liquid/solid ratio of 13.35 mL/g, 61.6 °C, 4.87 MPa, 100 vol % CO2, 50 min6.74 wt% CaO50.72
Ukwattage et al. [34]liquid/solid ratio of 5 mL/g, 40 °C, initial CO2 pressure of 6 MPa, 10 h39.8 wt% CaO, 7.3 wt% MgO7.66
Miao et al. [35]liquid/solid ratio of 10 mL/g, 5 kWh/m3 energy input in slurry, 60 °C, 15 vol % CO2, 2 h33.1 wt% CaO, 0.95 wt% MgO128
Montes-Hernandez et al. [36]liquid/solid ratio of 10 mL/g, 30 °C, initial CO2 pressure of 1 MPa, 18 h5.0 wt% CaO26.2
Bauer et al. [37]1500 rpm, liquid/solid ratio of 0.12 mL/g, 25–80 °C, initial CO2 pressure of 0.015 MPa, 2 h28.4 wt% Ca, 0.92 wt% Mg211
Ukwattage et al. [38]60 rpm, liquid/solid ratio of 3 mL/g, 60 °C, initial CO2 pressure of 3 MPa, 10 h24.8 wt% CaO, 13 wt% MgO27.1
Dananjayan et al. [39]900 rpm, liquid/solid ratio of 15 mL/g, 30 °C, initial CO2 pressure of 0.4 MPa, 2 h6.74 wt% CaO, 2.22 wt% MgO50.3
Patel et al. [40]liquid/solid ratio of 0.24 mL/g, 50 °C, 30 vol % CO237.25 wt% CaO, 0.45 wt% MgO40
Ho et al. [41]liquid/solid ratio of 100 mL/g, atmospheric pressure room temperature, 30 vol % CO2, 30 min3.44 wt% Ca, 0.82 wt% Mg16
Nyambura et al. [46]brine, 600 rpm, liquid/solid ratio of 2 mL/g, 30 °C, initial CO2 pressure of 4 MPa, 2 h9.2 wt% Ca, 2.44 wt% Mg71.8
Jo et al. [43]deionized water, liquid/solid ratio of 3 mL/g, 25 °C, 15 vol % CO2, 18 h7.2 wt% CaO, 1.5 wt% MgO19
1 M NH4 Cl, liquid/solid ratio of 3 mL/g, 25 °C, 15 vol % CO2, 18 h23
sea water, liquid/solid ratio of 3 mL/g, 25 °C, 15 vol % CO2, 18 h20
Ji et al. [42]0.5 M Na2CO3, liquid/solid ratio of 10 mL/g, 275 °C, initial CO2 pressure of 2 MPa, 2 h16.4 wt% CaO, 1.2 wt% MgO102
Ji et al. [44]0.5 mol/L piperazine solution, liquid/solid ratio of 10 mL/g, atmospheric pressure, 55 °C, 40 vol % CO2, 1.5 h24.3 wt% CaO, 0.9 wt% MgO102.9

2.2. Indirect Carbonation

Indirect mineral carbonation of coal fly ashes mainly involves the extraction of Ca2+ by acids or other solvents into an aqueous solution, followed by the carbonation reaction between the extracted Ca2+ and injected CO2. Indirect mineral carbonation allows for the production of pure carbonates due to the impurities (i.e., Si and Fe) that can be filtered out prior to carbonate precipitation [47].
He et al. [48,49] explored the effect of various extraction agents (i.e., NH4Cl, NH4NO3, and CH3COONH4) on Ca2+ extraction efficiency from COFA particles. The above three ammonia salts were all effective Ca2+ extraction agents, achieving a calcium leaching efficiency of about 35–40% within 1 h. CH4COONH4 was superior to the other two ammonia salts in promoting Ca2+ extraction of COFA. A carbonation efficiency of 41–47% was achieved when carbonating the leachate with CO2. Furthermore, a higher carbonation efficiency of ~90–93% can be obtained when NH4HCO3 is used instead of CO2 as the source of CO32−. By this method, 1 ton of fly ash could sequester 0.075 t of CO2, simultaneously producing 0.17 t of precipitated CaCO3. Jo et al. [50] investigated the effect of solid dosage, CaO content, CO2 flow rate, and solvent type on the extraction efficiency of Ca2+ from coal fly ashes. In fact, the extraction efficiency of Ca2+ was closely associated with the ultimate carbonation efficiency of fly ash. The results indicated that the Cn of fly ash was about 8 g-CO2/kg-fly ash using deionized water as the solvent at ambient temperature and pressure conditions with a solid dosage of 100 g/L and a CO2 flow rate of 2 mL/min.
Moreover, acetic acid was also used as the leachate to extract Ca2+ and Mg2+ from brown COFA to achieve indirect CO2 mineralization sequestration [51]. Most of the Ca and Mg within COFA could be dissolved into the solution at the initial stage of leaching, and the carbonation of the leached solution was conducted using a continuously stirred high-pressure autoclave. The CO2 was sequestered mainly in carbonate precipitates (i.e., CaCO3 and MgCO3) and water-soluble bicarbonate (i.e., Mg(HCO3)2). The carbonation of brown COFA-derived leachate exhibited a markedly lower global activation energy of 12.7 kJ/mol compared to that of natural minerals, contributing to maintaining a fast reaction rate. As shown in Figure 8, the maximum Cn of the fly ash was obtained at 60 ℃ for either CO2 being sequestrated in the form of the solid or liquid phase. Therefore, the maximum Cn could reach 264 g CO2/kg fly ash when considering both the contributions of carbonate precipitates and water-soluble bicarbonate. In addition, multiple-cycle leaching-carbonation and Mg2+ leaching kinetic modeling were studied during indirect carbonation of Victorian brown COFA for CO2 fixation [47]. It revealed that the extraction efficiency of Ca2+ and Mg2+ markedly decreased with the increased cycling number of ammonium chloride. The carbonation efficiency gradually dropped with the increase in ammonium chloride cycles as well.
An integrated circular system combining indirect carbonation of COFA with solution regeneration by bipolar membrane electrodialysis (BPED) was put forward by Ho et al. [52]. As shown in Figure 9, the system consisted of three main blocks, i.e., Ca2+ leaching from fly ash, CaCO3 precipitation, and acid and alkaline solution regeneration. The CO2 sequestration mechanism of COFA is described in Equation (2).
L e a c h a t e   s o l u t i o n C a 2 + + 2 N O 3 + C O 2   c a p t u r e   s o l u t i o n 2 N a + + C O 3 2 C a C O 3 ( s ) + 2 N a N O 3 ( a q )
BPED could regenerate the HNO3 and NaOH solutions (Equation (3)). The bipolar membranes could decompose the water into H+ and OH, and simultaneously dissociate the salt into Na+ and NO3 through electrodialysis, consequently regenerating the HNO3 and NaOH solutions.
N a N O 3 ( a q ) + H 2 O N a O H ( f o r   C O 2   c a p t u r e ) + H N O 3 ( f o r   l e a c h i n g )
The experiment results demonstrated that the Cn was ~11 g CO2/g fly ash, corresponding to a CO2 conversion of 92.9%. During the leaching stage, the adoption of a relatively lower nitric acid/calcium ratio was beneficial to the generation of a higher-purity CaCO3 precipitate.
A novel electrolytic carbonation system for the collaborative treatment of COFA, brine wastewater, and CO2 was proposed by Lu et al. [53]. As shown in Figure 10, the acidity produced by the electrolysis of brine electrolyte at the anode directly leached Ca2+ or Mg2+ from COFA. The Ca2+ or Mg2+ balanced the OH generated at the cathode to form Ca(OH)2 or Mg(OH)2, which then sequestered CO2 and simultaneously produced high-purity carbonate precipitates. The electrolysis contributed to the enhanced dissolution of fly ash, and then it increased the Cn from 9.75 to 18.42 g CO2/kg fly ash in the NaCl electrolyte.
In addition, protonated amine [54] and glycine [55,56] were used as reagents during the COFA-based leaching-mineralization course, which avoided the large consumption of exogenous acid-base reagents. The protonated amine could promote Ca2+ leaching from COFA, and simultaneously be converted into free amine [54]. Then, the free amine adsorbed the protons released by the precipitation reaction of CaCO3 and could achieve in-situ regeneration of protonated amine. The authors systematically investigated the CO2 mineralization performance of 13 typical amine-mediated COFA and the polymorph selection of CaCO3. Triethanolamine (TEA) exhibited the largest CaCO3 yield of 56.8%, and the corresponding Ca utilization efficiency was 16.8%. Moreover, the amines had a pronounced control effect on the CaCO3 product’s size, polymorph, and morphology. As shown in Figure 11, the primary amino group was superior to the secondary and tertiary ones in promoting the formation of vaterite. The introduction of a side chain was prone to make the vaterite transform to calcite, and the grain size of the vaterite was inversely proportional to the length of the side chain.
To achieve CO2 sequestration and CaCO3 production, Zheng et al. [55,56] designed a glycine-mediated leaching-mineralization method for COFA. The process can simultaneously achieve the desirable CO2 sequestration efficiency of COFA and produce high-purity CaCO3 products in an in-situ recyclable glycine solution. After leaching for 1 h, a mineralization efficiency of 74.4% and a maximum CaCO3 yield of 98.8 g/kg fly ash were achieved in a 2 M glycine solution with a fly ash dosage of 200 g/L. The reaction mechanism of the glycine-mediated leaching-mineralization approach through glycine and COFA (COFA) was illustrated in Figure 12. During the leaching stage, Gly0 could act as the proton donor and chelating agent that markedly promoted Ca2+ leaching, simultaneously producing Gly. As an excellent CO2 absorbent, Gly accelerates the mass transfer of gaseous CO2 to the liquid phase by offering extra CO2 reaction pathways. The carboxyl of Gly could also bind to the protons derived from carbonation reactions in the mineralization step, acting as a proton receptor. Moreover, the Gly species acted as a desirable crystal regulator in carbonate precipitation, which contributed to extending the continuous conversion and growth of CaCO3 crystals.
An integrated system coupling mineral carbonation with direct air capture under atmospheric conditions was put forward by Ragipani et al. [57], as depicted in Figure 13. The evenly mixed slurries of COFA and concentrated alkali carbonate were used as recyclable solvents. Sodium hydroxide solvent was used to sequestrate CO2 (Equation (4)) during the DAC process. The Equations (5) and (6) described the reaction generating NaOH solution via the carbonation between the COFA and Na2CO3. conditions. It was found that the coupled system achieved a high CaCO3 conversion of ~80% in a 1.9 M Na2CO3 solution for 1 h. Moreover, the cost of sequestering 1 ton of CO2 was about $116−133, while the process based on life cycle assessment emitted 0.03−0.25 t-CO2e/t-CO2-sequestered.
2 N a O H ( a q ) + C O 2 ( g ) N a 2 C O 3 ( a q ) + H 2 O
2 C a O . S i O 2 ( s ) + 2 N a 2 C O 3 ( a q ) + 2 H 2 O 2 C a C O 3 ( s ) + S i O 2 ( s ) + 4 N a O H ( a q )
C a O . A l 2 O 3 . 2 S i O 2 ( s ) + N a 2 C O 3 ( a q ) + 4 H 2 O C a C O 3 ( s ) + 2 A l ( O H ) 3 + 2 S i O 2 ( s ) + 2 N a O H ( a q )

3. CO2 Mineralization in Conjunction with Fly Ash-Derived by-Products Utilization

As mentioned above, the carbonation of COFAes can effectively achieve the fixation of CO2. In addition, the further utilization of fly ash-derived by-products has also received extensive attention from researchers. Ren et al. [58] investigated the feasibility of combining the preparation of composite gravel with CO2 absorption using COFA. The composite gravel pellets were prepared from mixtures of COFA, gypsum, Portland Cement and water, and the pellets were subjected to carbonation curing using flue gas. Figure 14a illustrates the prepared composite gravel pellets. The composite gravel pellets contained a 20 wt% cement content and 1:11(A11) ratio of gypsum to COFA, exhibiting the highest compressive strength of 11.11 MPa at the age of 28 days, as shown in Figure 14b. It was positively correlated with its CO2 adsorption capacity, and the CO2 adsorption for composite gravel pellets reached 4.789% with a 1:11 ratio of gypsum to COFA (Figure 14c).
Moreover, a new approach of employing the carbonated slurry of coal fly ashes during the CO2 sequestration process to inhibit spontaneous combustion of coal (SCC) in the coal mining process was put forward [20]. The COFA reached a Cn of 54.9 g CO2/kg fly ash with a liquid/solid ratio of 6 at 2 MPa. The carbonated fly ash slurry demonstrated a superior capability to inhibit SCC compared to the uncarbonated slurry. The temperature increase at the crossing point reached 13.8 °C in comparison to the origin coal when the carbonated fly ash slurry-to-coal ratio was set at 3:1. It greatly diminished the risk of SCC, indicating carbonated fly ash slurry possessing an apparent suppression function on SCC. The carbonated coal fly ashes were used as the mineral admixture for concrete or cement mortar [59,60]. Chen et al. [60] studied the property of cement mortar containing carbonated coal fly ashes and the combined impact of carbonation curing. It showed that carbonation treatment decreased the proportion of free CaO and the hydration heat of coal fly ashes. Therefore, the expansion of mortar specimens incorporating carbonated fly ash was greatly mitigated in comparison to that of specimens using fresh fly ash [59]. In addition, Freire et al. [61] produced geopolymers with COFA and rice husk ash, which were adopted as the CO2 capture materials, and various geopolymer formulations were tried. The calcined rice husk ash activated via NaOH was the best precursor to prepare geopolymer that exhibited superior CO2 capture performance.
The dual needs of CO2 emission reduction and COFA treatment require the design of an innovative pathway to simultaneously achieve CO2 sequestration and further conversion into useful carbonates. Yin et al. [62] conducted a range of experiments using fly ash and a mixture of fly ash and bottom ash to study CO2 mineralization function with regenerable solvents (i.e., 2.5 M sodium glycinate and 30 wt% MEA). A mixture of fly ash and bottom ash was superior to sole fly ash, which obtained the highest CO2 mineralization degree of non-calcium carbonate content. Moreover, nanoscale calcium carbonate could be successfully produced from dissolved calcium through CO2-loaded sodium glycinate and CTAB surfactant. The detailed reaction mechanism for the integrated approach of CO2 mineralization and nanoscale calcium carbonate production using COFA was demonstrated in Figure 15. Therefore, the route of directly sequestering CO2 in COFA and simultaneously producing nanoscale CaCO3 with regenerable CO2 capture solvents was feasible.
In addition, Monasterio-Guillot et al. [63] were the first to explore the combined effects of CO2 sequestration, zeolite production, and simultaneous heavy metal element trapping at the carbonation stage of COFA. The carbonation efficiency was as high as 79% (a net Cn of 45 g CO2/kg fly ash) c under mild hydrothermal conditions for Class F fly ash (3.72 wt% CaO, 1.74 wt% MgO). Simultaneously, the amorphous precursors of zeolite and different crystalline zeolites with a yield as high as 60 wt% were produced. The potentially toxic elements within COFA were effectively solidified in the newly generated calcite and zeolite products, avoiding the leaching of toxic elements.
As illustrated in Figure 16, a green and facile pathway to synthesize mesoporous γ-Al2O3 from COFA while simultaneously achieving on-site CO2 utilization was investigated by Yan et al. [64]. The lime-sinter route was used to extract aluminum from COFA. The COFA mixed with CaCO3 was calcined at 1390 °C and then the calcined products were dissolved in Na2CO3 solution at 70 °C, achieving a high aluminum extraction efficiency of 87.42%. Then, the flue gas containing 15 vol% CO2 and 85 vol% N2 after being purified was injected into the extracted liquid to achieve CO2-assisted precipitation of Al(OH)3 precursors (Equation (7)). Ultimately, the mesoporous γ-Al2O3 (high surface area of 230.3 m2.g−1) was produced by calcining the Al(OH)3 at 400 °C or 550 °C (Equation (8)). This proposed pathway appeared to be desirable for scaled-up production of mesoporous γ-Al2O3 by integrating the on-site recycling of COFA and CO2 utilization.
2 N a A l O 2 ( a q ) + C O 2 ( g ) + 3 H 2 O N a 2 C O 3 ( a q ) + 2 A l ( O H ) 3 ( s )
2 A l ( O H ) 3 ( s ) A l 2 O 3 ( s ) + 3 H 2 O

4. CO2 Sequestration of COFA by Dry Carbonation Process

In gas-solid carbonation, CO2 reacts with free CaO within COFA to generate CaCO3. Patel et al. [40] studied the CO2 sequestration performance of COFA with reaction temperatures ranging from 500°C to 750 °C. Figure 17 demonstrated that the CO2 sequestration level apparently increased with the temperature being leveled up to 700 °C in the initial stage. However, the CO2 sequestration efficiency gradually dropped to below ~50% after the reaction time exceeding 10 min. Moreover, an apparent reduction in CO2 sequestration efficiency was observed in the rapid reaction regime when the temperature reached 750 °C due to the simultaneous partial decomposition of CaCO3.
Revathy et al. [33] designed the carbonation experiments of (Indian COFA (class F type-CaO < 10 wt%) via the gas-solid route using response surface methodology, and temperature and pressure were the main factors studied. The interaction between the factors (i.e., temperature and pressure) and their impacts on response (the reduction of CO2 concentration) was analyzed. The temperature affected the carbonation efficiency of COFA because of the remarkable influence of temperature change on CO2 concentration. At a certain temperature, the pressure played a profound role in the gas-solid carbonation of COFA. It was noted that the influence of temperature seemed to be more remarkable compared to that of pressure on the carbonation of fly ash when both temperature and pressure were changed. The global optimum solution was identified as a temperature of 44.86 °C and pressure of 24 bar, obtaining a reduction in CO2 concentration of 18.2%. The COFA exhibited a maximum Cn of 20.0 g CO2/kg fly ash under gas-solid carbonation.
Liu et al. [65] studied the influence of temperature, CO2 concentration, steam concentration, and reaction duration on the CO2 fixation performance of COFA via a direct gas-solid carbonation approach. It was found that the increased temperature and the concentration of CO2 and H2O(g) contributed to the improved CO2 fixation efficiency. The effects of temperature and H2O(g) were more apparent than that of CO2 content. The COFA achieved a maximum Cn of 60 g CO2/kg fly ash corresponding to a fixation efficiency of 28.74% at 600 oC with 20% H2O(g) addition. Moreover, Ćwik et al. [66] carried out a mineral carbonation experiment of high-calcium COFA using a continuous flow reactor with different temperatures and CO2 pressures. The dry and moist conditions were comparatively investigated. Results exhibited that COFA achieved a maximal Cn of 117.7 g CO2/kg fly ash (corresponding to a carbonation efficiency of 48.14%) under gas-solid conditions. Moderate amounts of H2O(g) contained in the CO2 gas flow were beneficial to the improved carbonation efficiency. The promotion effect of steam resulted from the formation of Ca(OH)2 or transient Ca(OH)2 and the enhanced CO2 molecular mobility at the carbonation stage [67,68].
Moreover, the CO2 mineral carbonation of COFA doped with carbide slag was conducted in a pilot-scale circulating entrained-flow bed (CEB) rector (Figure 18). In a semi-dry atmosphere, the carbonation efficiency was enhanced 4 times higher than that of carbide-slag-doped COFA. It was mainly attributed to a thin liquid film formed on the surface of fly ash that facilitated CO2 diffusion. At 550 °C, a maximal carbonation efficiency of 55% was obtained when adding 15% steam to the feed and injecting 6 kg/h of water into the rector system [69].

5. CO2 Sequestration Materials Derived from COFA

Usually, COFA is rich in silica and alumina, which is beneficial to enhance the high-temperature CO2 capture stability of CaO-based sorbents. Chen et al. [18,70] spared efforts to seeking a potentially effective approach to improve the capability of CaO-based sorbents mixed with COFA for CO2 capture. The comparative studies on the effect of pretreatments (i.e., grinding, calcination, hydration, and leaching) of COFA on CO2 capture performance of fly ash-modified CaO-based sorbents were conducted. The pretreatments of grinding and calcination for fly ash contributed to the improved CO2 uptake capability for CaO-based sorbents. Acidic conditions were superior to basic conditions in enhancing CO2 capture capacity and reactivity for CaO-based sorbents hydrated with fly ash [70]. The improved pore structure and the formation of inert stabilizers (i.e., Ca12Al14O33 and CaSiO3) within the CaO-based sorbents are the main reasons [71,72,73]. Particularly, the fly ash-modified sorbents with a mass ratio of slag/calcined calcium carbonate of 1:2, exhibited the CO2 capture capacity of 0.19 g CO2/g sorbent in the 30th cycle under severe calcination conditions, which was 3.8 times that of pure CaO [18].
Yan et al. [17] developed a green approach that involved silicon extraction from COFA by the method of alkali dissolution. Then, the synthetic nano silica was incorporated into CaO-based sorbents to improve both their cyclic CO2 capacities and their sorption kinetics. Results exhibited that the synthetic nano silica-stabilized CaO-based sorbent still showed a higher CO2 uptake of 0.2 g CO2/g sorbent in short carbonation duration after 30 cycles (Figure 19), apparently increased by 155% over pure CaO. What was more important, the synthetic nano silica-stabilized CaO-based sorbent achieved approximately 90% of the total CO2 uptake within ~20 s. The extremely fast CO2 capture rate was very key for practical applications. The superior CO2 capture performance of synthetic nano silica-stabilized sorbent was mainly ascribed to the evenly distributed Ca2SiO4, yielding a large amount of small pores directly exposed to CO2 during cycles.
The fly ash-stabilized, CaO-based sorbents were prepared via a physically dry-mixing method with different CaO precursors (i.e., CaO, Ca(OH)2, CaCO3, and CaC2O4.H2O) [74]. After 30 cycles, the sorbent synthesized from CaC2O4.H2O (90 wt% of CaO in the sorbent) exhibited the highest CO2 uptake of 0.38 g CO2/g CaO. This sorbent still showed a desirable capacity of 0.27 g CO2/g CaO even after being calcined at 920 °C in pure CO2. The excellent performance of the fly ash-stabilized, CaO-based sorbents for cyclic CO2 capture mainly resulted from the formation of highly dispersed inert Ca2Al2SiO7 within the sorbent. Sreenivasulu et al. [75] further conducted the thermodynamic and kinetic studies of CO2 capture using a COFA-doped sorbent (50 wt% CaO, 10 wt% MgO, and 40 wt% COFA). The thermodynamic studies confirmed both the feasibility of the reaction and the positive role of COFA in improving CO2 capture and reducing regeneration temperatures. They also proposed a model for both the reaction and diffusion kinetic control stages, which had been verified through the experimental data obtained. It was found that the rate constants of 2.5 and 0.8 min−1 and activation energies of 23 and 30 kJ/mol were evaluated for reaction- and diffusion-controlled stages at 650 °C, respectively. The results clearly indicated the improved rate achieved by COFA-doped CaO-based sorbents along with their thermal stability.
Moreover, the CO2 uptake capacity of hydrated lime mixed with fly ash under low-temperature and humid conditions was also investigated [76]. The reaction condition was set at 60 °C with 70% relative humidity. The fly ash-doped CaO-based sorbents containing <50 wt% of fly ash exhibited higher CO2 uptake than hydrated lime without doping fly ash (~0.15 g CO2/g sorbent). The sorbent doped with 30 wt% of fly ash showed a particularly high maximum CO2 uptake of 0.26 g of CO2/g sorbent. The enhanced microstructure resulting from the generation of calcium silicate hydrates was responsible for the improved reactivities of the fly ash-doped sorbents.
Moreover, various kinds of zeolites synthesized from COFAes were used to capture CO2 in the flue gas. As shown in Figure 20, the typical octahedral prism and cube-shaped crystals were observed for the synthetic zeolites derived from COFAes, indicating the good generation of the desired zeolitic materials. Aquino et al. [77] comparatively measured the adsorption capacity of two synthetic zeolites (i.e., NaX and NaA types) from COFA and commercial zeolites and further assessed their performance in the temperature swing adsorption (TSA) testing process. It was found that NaX and NaA-type zeolites exhibited CO2 adsorption capacities of 1.97 and 1.37 mmol CO2/g adsorbent at 303 K, which were close to those of commercial zeolites. More importantly, the synthetic zeolites showed highly stable adsorption capacity during five cycles, demonstrating the possibility of practical application in TSA processes.
Verrecchia et al. [14] focused on studying the effect of NaOH/COFA mass ratio, crystallization temperature, and duration on the synthesis of COFA-derived zeolites by the combined approach of fusion and hydrothermal. The major products contained NaX and amorphous compounds that were produced at 90 °C with a NaOH/COFA ratio of 1.2 for 7 h. They exhibited a capacity of 2.18 mol CO2/kg adsorbent for CO2 adsorption, accounting for 57% of the capacity of commercial 13X zeolites. Further, the central composite full factorial method was used to design experiments to optimize the synthesis parameters, aiming to improve the capacity of COFA-derived zeolites for CO2 adsorption. It indicated that the synthetic zeolite produced with the experimental conditions of 1.4 NaOH/COFA, 80 °C, and 7 h exhibited a markedly increased adsorption capacity of 3.3 mol CO2/kg adsorbent (~86% of commercial 13X zeolites). Muriithi et al. [15] also successfully synthesized two types of zeolites (NaX and NaA type) from COFAes used for CO2 capture. The synthetic NaX-type zeolite possessed hierarchical morphology with platy structures arranged in spherical clusters and exhibited desirable CO2 physisorption properties.
Moreover, the collaborative activation method of microwave and ultrasound was adopted to synthesize K-MER zeolites with desirable crystallinity and high purity [78]. Comprehensively adjusting the crystallization duration, pressures, and dynamic pressure programs could produce K-MER zeolites exhibiting various morphologies. At 25 °C, the as-synthetic K-MER zeolites exhibited a maximal adsorption volume of CO2 of 47.58 cm3/g. The as-synthetic zeolites could also be used as catalysts to promote the cyanoethylatic reaction between methanol and acrylonitrile, which achieved a high acrylonitrile conversion rate of 99 wt% for 0.5 h under a pre-pressure of 1.5 MPa.
Furthermore, the COFA could be adopted as the raw material for producing synthetic hierarchical porous nano silica [16], nanoparticles MCM-41 [79], calcium silicate hydrate support [80], and silica-alumina aerogel [19] for CO2 captured adsorbents. The hierarchical porous nano silica was mainly prepared through the collaborative method of microwave-assisted alkaline extraction and hydrothermal synthesis from fly ash derived from coal gasification [16]. The hierarchical porous nano silica had the isosteric adsorption heat of −30.06 kJ/mol, demonstrating that CO2 adsorption involved both physisorption and chemisorption. The hierarchical porous nano silica exhibited a CO2 adsorption capacity of 1.63 mmol CO2/g adsorbent at 30 °C and 1 bar. The nanoparticle MCM-41 was functionalized with (3-aminopropyl) triethoxysilane and dispersed in diethylenetriamine (DETA) solution through ultrasonic dispersion method to prepare nanofluids for CO2 capture [79]. The nanofluids coupled with high gravity technology markedly enhanced the CO2 capture efficiency because of the improvement of grazing and hydrodynamic effect in a high gravity field (as shown in Figure 21), reaching ~95.7%. Qu et al. [80] synthesized highly porous COFA-derived calcium silicate hydrate (CSH) using the pore-expanded assistance of azeotropic distillation. The COFA-derived calcium silicate hydrate was then used as the support to prepare PEI@CSH adsorbents via impregnation. The 60% [email protected] adsorbent exhibited the excellent CO2 uptake of 198 mg CO2/g adsorbent under ideal regeneration conditions, and after 10 cycles, the adsorption capacity remained at 103 mg CO2/g adsorbent even under severe regeneration conditions.
The silica-alumina aerogel derived from COFA possessed good microstructure, which could also be used as the inert support to prepare K2CO3-based CO2 adsorbent with a wetness impregnation method [19]. The synthetic silica-alumina aerogel support exhibited a relatively high specific surface area of 400 m2/g and a specific pore volume of 1.9 cm3/g, and the proportion of mesopores within the support was over 99%. Therefore, the K2CO3-based adsorbent with K2CO3 load of 30 wt% achieved the maximal adsorption capacity of 2.02 mmol CO2/g adsorbent under the conditions of 60 °C, 15 vol% water vapor, and 15 vol% CO2.

6. Conclusions and Prospects

This paper summarizes the CO2 sequestration technologies using COFA reported in the literature. The relevant technical routes and representative research results from previous studies in this field are discussed and analyzed. The mineral CO2 sequestration using COFA is one of the most promising technologies, which can be classified into direct and indirect carbonation. The direct CO2 mineralization route is simple and highly practicable, which eliminates the extraction step of reactive components and minimizes the consumption of chemical reagents. Indirect CO2 mineralization involves the extraction of reactive Ca2+ from COFA with leaching agents (i.e., acid, ammonium salt, etc.) followed by carbonation reactions, which allow the production of high-value pure carbonates.
Moreover, the by-products derived from the CO2 mineralization using COFA can be further utilized as feedstock to prepare composite gravel, cement mortar, geopolymers, mesoporous γ-Al2O3, and zeolite for toxic element trapping. The free CaO within COFA makes it possible to sequester CO2 by dry process, and steam addition can effectively enhance carbonation efficiency. In addition, the silica and alumina contained in COFA can be used as materials for the synthesis of highly efficient CO2 adsorbents. The COFA-modified CaO-based sorbents exhibit highly stable CO2 capture performance due to the formation of extensive inert stabilizers, such as Ca12Al14O33, Ca2Al2SiO7, CaSiO3, and Ca2SiO4. The CO2 adsorption capability of the synthetic zeolites (i.e., NaX, NaA, and K-MER type) from coal fly ashes is comparable to that of commercial zeolites. Moreover, the nanoparticles MCM-41, calcium silicate hydrate, and silica-alumina aerogel derived from COFA can be used as the porous support of amine- or K2CO3-based adsorbents for low-temperature CO2 capture.
The research on the development of CO2 sequestration technologies using COFA in the past 10 years is summarized and analyzed. There are still some issues to be studied and solved.
(1)
Although extensive research has been carried out to enhance the CO2 mineralization performance of COFA, the current research is limited to the optimization of experimental parameters at the laboratory scale. In the future, this technology needs to be promoted for industry, which needs to consider the cost of various raw materials and energy consumption. Moreover, the carbon assets obtained from CO2 fixation and the additional economic benefits of valued carbonation by-products also need to be considered. Therefore, it is urgent to achieve an integrated economic feasibility analysis or life cycle assessment concerning the industrial-scale CO2 mineralization of COFA. The researchers can obtain the technical cost of CO2 sequestration for direct and indirect carbonation of COFA via the above technical process economic analysis. This will be able to better demonstrate the practicality of the direct and indirect carbonation of COFA technology and the potential for industrial-scale promotion.
(2)
COFA contains major associated elevated soluble trace elements (e.g., As, Cr, Mn, Cu, Sr, Ce, and V, etc.) that are potentially toxic to the biological system [63,81,82]. The heavy metal toxic elements are released into the surrounding environment (adversely affecting the plant and soil quality) by leaching of COFA once these are ponded or landfilled. Furthermore, these potentially toxic trace elements can reenter the food chain and human life cycle from these disposal sites via certain pathways [83]. Relevant scholars have investigated the accelerated carbonation of phosphogypsum [84] and municipal solid waste incineration (MSWI) fly ash [21,81,85], and found that the carbonation reaction can solidify the heavy metal elements within waste to a certain extent in the realization of CO2 sequestration. Hence, it is necessary to investigate the leaching characteristics of potentially toxic elements in COFA during the wet CO2 mineralization process.
(3)
For the synthesis of zeolites for CO2 capture, the types of zeolite products are not rich enough. More technologies and processes should be developed to prepare low-cost and high-grade zeolites. Furthermore, the development of zeolites with highly selective adsorption and multi-effect functions should be the focus of research.

Author Contributions

Conceptualization, data curation, and writing—original draft preparation, L.J.; validation, formal analysis and visualization, L.C. and Y.Z.; supervision, G.L.; writing—review and editing, funding acquisition, J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Jiangsu Education Department Fund (23KJB470025) and Graduate Research and Innovation Projects of Jiangsu Province (SJCX23_0605).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wee, J.-H. A review on carbon dioxide capture and storage technology using coal fly ash. Appl. Energy 2013, 106, 143–151. [Google Scholar] [CrossRef]
  2. Dindi, A.; Quang, D.V.; Vega, L.F.; Nashef, E.; Abu-Zahra, M.R. Applications of fly ash for CO2 capture, utilization, and storage. J. CO2 Utilizat. 2019, 29, 82–102. [Google Scholar] [CrossRef]
  3. Wang, C.; Xu, G.; Gu, X.; Gao, Y.; Zhao, P. High value-added applications of coal fly ash in the form of porous materials: A review. Ceram. Int. 2021, 47, 22302–22315. [Google Scholar] [CrossRef]
  4. Łach, M.; Grela, A.; Pławecka, K.; Guigou, M.D.; Mikuła, J.; Komar, N.; Bajda, T.; Korniejenko, K. Surface modification of synthetic zeolites with Ca and HDTMA compounds with determination of their phytoavailability and comparison of CEC and AEC parameters. Materials 2022, 15, 4083. [Google Scholar] [CrossRef]
  5. Wang, J.; Yang, Y.; Jia, Q.; Shi, Y.; Guan, Q.; Yang, N.; Wang, Q.; Ning, P. A critical review on solid waste derived CO2 capturing materials. ChemSusChem 2019, 15, 4083. [Google Scholar]
  6. Sun, J.; Liang, C.; Tong, X.; Guo, Y.; Li, W.; Zhao, C.; Zhang, J.; Lu, P. Evaluation of high-temperature CO2 capture performance of cellulose-templated CaO-based pellets. Fuel 2019, 239, 1046–1054. [Google Scholar] [CrossRef]
  7. Sheng, C.; Li, Y. Experimental study of ash formation during pulverized coal combustion in O2/CO2 mixtures. Fuel 2008, 87, 1297–1305. [Google Scholar] [CrossRef]
  8. Xu, M.; Yu, D.; Yao, H.; Liu, X.; Qiao, Y. Coal combustion-generated aerosols: Formation and properties. Proc. Combust. Inst. 2011, 33, 1681–1697. [Google Scholar] [CrossRef]
  9. Uliasz-Bocheńczyk, A.; Pawluk, A.; Pyzalski, M. The mineral sequestration of CO2 with the use of fly ash from the co-combustion of coal and biomass. Gospod. Surowc. Mineral. 2017, 33, 143–155. [Google Scholar] [CrossRef]
  10. Jia, B.; Tsau, J.-S.; Barati, R. A review of the current progress of CO2 injection EOR and carbon storage in shale oil reservoirs. Fuel 2019, 236, 404–427. [Google Scholar] [CrossRef]
  11. Jia, B.; Xian, C.; Tsau, J.-S.; Zuo, X.; Jia, W. Status and outlook of oil field chemistry-assisted analysis during the energy transition period. Energ. Fuel 2022, 36, 12917–12945. [Google Scholar] [CrossRef]
  12. Bui, M.; Adjiman, C.S.; Bardow, A.; Anthony, E.J.; Boston, A.; Brown, S.; Fennell, P.S.; Fuss, S.; Galindo, A.; Hackett, L.A. Carbon capture and storage (CCS): The way forward. Energy Environ. Sci. 2018, 11, 1062–1176. [Google Scholar] [CrossRef]
  13. Bobicki, E.R.; Liu, Q.; Xu, Z.; Zeng, H. Carbon capture and storage using alkaline industrial wastes. Prog. Energy Combust. 2012, 38, 302–320. [Google Scholar] [CrossRef]
  14. Verrecchia, G.; Cafiero, L.; de Caprariis, B.; Dell’Era, A.; Pettiti, I.; Tuffi, R.; Scarsella, M. Study of the parameters of zeolites synthesis from coal fly ash in order to optimize their CO2 adsorption. Fuel 2020, 276, 118041. [Google Scholar] [CrossRef]
  15. Muriithi, G.N.; Petrik, L.F.; Doucet, F.J. Synthesis, characterisation and CO2 adsorption potential of NaA and NaX zeolites and hydrotalcite obtained from the same coal fly ash. J. CO2 Utilizat. 2020, 36, 220–230. [Google Scholar] [CrossRef]
  16. Liu, Y.; Wang, Z.; Zhao, W.; Hou, J.; Cui, L.; Zou, L.; Li, C.; Li, H.; Wu, Y.; Xu, R. Hierarchical porous nanosilica derived from coal gasification fly ash with excellent CO2 adsorption performance. Chem. Eng. J. 2023, 455, 140622. [Google Scholar] [CrossRef]
  17. Yan, F.; Jiang, J.; Li, K.; Liu, N.; Chen, X.; Gao, Y.; Tian, S. Green synthesis of nanosilica from coal fly ash and its stabilizing effect on CaO sorbents for CO2 capture. Environ. Sci. Technol. 2017, 51, 7606–7615. [Google Scholar] [CrossRef]
  18. Chen, H.; Khalili, N. Fly-ash-modified calcium-based sorbents tailored to CO2 capture. Ind. Eng. Chem. Res. 2017, 56, 1888–1894. [Google Scholar] [CrossRef]
  19. Guo, B.; Zhang, J.; Wang, Y.; Qiao, X.; Xiang, J.; Jin, Y. Study on CO2 adsorption capacity and kinetic mechanism of CO2 adsorbent prepared from fly ash. Energy 2023, 263, 125764. [Google Scholar] [CrossRef]
  20. Shao, X.; Qin, B.; Shi, Q.; Yang, Y.; Ma, Z.; Xu, Y.; Hao, M.; Jiang, Z.; Jiang, W. Study on the sequestration capacity of fly ash on CO2 and employing the product to prevent spontaneous combustion of coal. Fuel 2023, 334, 126378. [Google Scholar] [CrossRef]
  21. Bertos, M.F.; Li, X.; Simons, S.; Hills, C.; Carey, P. Investigation of accelerated carbonation for the stabilisation of MSW incinerator ashes and the sequestration of CO2. Green Chem. 2004, 6, 428–436. [Google Scholar] [CrossRef]
  22. Muriithi, G.N.; Petrik, L.F.; Fatoba, O.; Gitari, W.M.; Doucet, F.J.; Nel, J.; Nyale, S.M.; Chuks, P.E. Comparison of CO2 capture by ex-situ accelerated carbonation and in in-situ naturally weathered coal fly ash. J. Environ. Manag. 2013, 127, 212–220. [Google Scholar] [CrossRef]
  23. Ji, L.; Yu, H.; Zhang, R.; French, D.; Grigore, M.; Yu, B.; Wang, X.; Yu, J.; Zhao, S. Effects of fly ash properties on carbonation efficiency in CO2 mineralisation. Fuel Process. Technol. 2019, 188, 79–88. [Google Scholar] [CrossRef]
  24. Yuan, Q.; Yang, G.; Zhang, Y.; Wang, T.; Wang, J.; Romero, C.E. Supercritical CO2 coupled with mechanical force to enhance carbonation of fly ash and heavy metal solidification. Fuel 2022, 315, 123154. [Google Scholar] [CrossRef]
  25. Yuan, Q.; Zhang, Y.; Wang, T.; Wang, J.; Romero, C.E. Mineralization characteristics of coal fly ash in the transition from non-supercritical CO2 to supercritical CO2. Fuel 2022, 318, 123636. [Google Scholar] [CrossRef]
  26. Baciocchi, R.; Costa, G.; Lategano, E.; Marini, C.; Polettini, A.; Pomi, R.; Postorino, P.; Rocca, S. Accelerated carbonation of different size fractions of bottom ash from RDF incineration. Waste Manag. 2010, 30, 1310–1317. [Google Scholar] [CrossRef] [PubMed]
  27. Shih, S.-M.; Lin, J.-P.; Shiau, G.-Y. Dissolution rates of limestones of different sources. J. Hazard. Mater. 2000, 79, 159–171. [Google Scholar] [CrossRef] [PubMed]
  28. Sanchez, F.; Gervais, C.; Garrabrants, A.; Barna, R.; Kosson, D. Leaching of inorganic contaminants from cement-based waste materials as a result of carbonation during intermittent wetting. Waste Manag. 2002, 22, 249–260. [Google Scholar] [CrossRef] [PubMed]
  29. Siriwardena, D.P.; Peethamparan, S. Quantification of CO2 sequestration capacity and carbonation rate of alkaline industrial byproducts. Constr. Build. Mater. 2015, 91, 216–224. [Google Scholar] [CrossRef]
  30. Back, M.; Kuehn, M.; Stanjek, H.; Peiffer, S. Reactivity of alkaline lignite fly ashes towards CO2 in water. Environ. Sci. Technol. 2008, 42, 4520–4526. [Google Scholar] [CrossRef]
  31. La Plante, E.C.; Mehdipour, I.; Shortt, I.; Yang, K.; Simonetti, D.; Bauchy, M.; Sant, G.N. Controls on CO2 mineralization using natural and industrial alkaline solids under ambient conditions. ACS Sustain. Chem. Eng. 2021, 9, 10727–10739. [Google Scholar] [CrossRef]
  32. Pan, S.-Y.; Hung, C.-H.; Chan, Y.-W.; Kim, H.; Li, P.; Chiang, P.-C. Integrated CO2 fixation, waste stabilization, and product utilization via high-gravity carbonation process exemplified by circular fluidized bed fly ash. ACS Sustain. Chem. Eng. 2016, 4, 3045–3052. [Google Scholar] [CrossRef]
  33. Revathy, T.D.R.; Ramachandran, A.; Palanivelu, K. Carbon capture and storage using coal fly ash with flue gas. Clean Technol. Environ. Policy 2021, 24, 1053–1071. [Google Scholar] [CrossRef]
  34. Ukwattage, N.; Ranjith, P.; Wang, S. Investigation of the potential of coal combustion fly ash for mineral sequestration of CO2 by accelerated carbonation. Energy 2013, 52, 230–236. [Google Scholar] [CrossRef]
  35. Miao, E.; Du, Y.; Zheng, X.; Zhang, X.; Xiong, Z.; Zhao, Y.; Zhang, J. Kinetic analysis on CO2 sequestration from flue gas through direct aqueous mineral carbonation of circulating fluidized bed combustion fly ash. Fuel 2023, 342, 127851. [Google Scholar] [CrossRef]
  36. Montes-Hernandez, G.; Perez-Lopez, R.; Renard, F.; Nieto, J.; Charlet, L. Mineral sequestration of CO2 by aqueous carbonation of coal combustion fly-ash. J. Hazard. Mater. 2009, 161, 1347–1354. [Google Scholar] [CrossRef]
  37. Bauer, M.; Gassen, N.; Stanjek, H.; Peiffer, S. Carbonation of lignite fly ash at ambient T and P in a semi-dry reaction system for CO2 sequestration. Appl. Geochem. 2011, 26, 1502–1512. [Google Scholar] [CrossRef]
  38. Ukwattage, N.L.; Ranjith, P.; Yellishetty, M.; Bui, H.H.; Xu, T. A laboratory-scale study of the aqueous mineral carbonation of coal fly ash for CO2 sequestration. J. Clean. Prod. 2015, 103, 665–674. [Google Scholar] [CrossRef]
  39. Dananjayan, R.R.T.; Kandasamy, P.; Andimuthu, R. Direct mineral carbonation of coal fly ash for CO2 sequestration. J. Clean. Prod. 2016, 112, 4173–4182. [Google Scholar] [CrossRef]
  40. Patel, A.; Basu, P.; Acharya, B. An investigation into partial capture of CO2 released from a large coal/petcoke fired circulating fluidized bed boiler with limestone injection using its fly and bottom ash. J. Environ. Chem. Eng. 2017, 5, 667–678. [Google Scholar] [CrossRef]
  41. Ho, H.-J.; Iizuka, A.; Shibata, E. Utilization of low-calcium fly ash via direct aqueous carbonation with a low-energy input: Determination of carbonation reaction and evaluation of the potential for CO2 sequestration and utilization. J. Environ. Manag. 2021, 288, 112411. [Google Scholar] [CrossRef] [PubMed]
  42. Ji, L.; Yu, H.; Wang, X.; Grigore, M.; French, D.; Gözükara, Y.M.; Yu, J.; Zeng, M. CO2 sequestration by direct mineralisation using fly ash from Chinese Shenfu coal. Fuel Process. Technol. 2017, 156, 429–437. [Google Scholar] [CrossRef]
  43. Jo, H.Y.; Ahn, J.-H.; Jo, H. Evaluation of the CO2 sequestration capacity for coal fly ash using a flow-through column reactor under ambient conditions. J. Hazard. Mater. 2012, 241, 127–136. [Google Scholar] [CrossRef]
  44. Ji, L.; Zheng, X.; Zhang, L.; Feng, L.; Li, K.; Yu, H.; Yan, S. Feasibility and mechanism of an amine-looping process for efficient CO2 mineralization using alkaline ashes. Chem. Eng. J. 2022, 430, 133118. [Google Scholar] [CrossRef]
  45. Ji, L.; Yu, H.; Li, K.; Yu, B.; Grigore, M.; Yang, Q.; Wang, X.; Chen, Z.; Zeng, M.; Zhao, S. Integrated absorption-mineralisation for low-energy CO2 capture and sequestration. Appl. Energy 2018, 225, 356–366. [Google Scholar] [CrossRef]
  46. Nyambura, M.G.; Mugera, G.W.; Felicia, P.L.; Gathura, N.P. Carbonation of brine impacted fractionated coal fly ash: Implications for CO2 sequestration. J. Environ. Manag. 2011, 92, 655–664. [Google Scholar] [CrossRef]
  47. Hosseini, T.; Selomulya, C.; Haque, N.; Zhang, L. Indirect carbonation of victorian brown coal fly ash for CO2 sequestration: Multiple-cycle leaching-carbonation and magnesium leaching kinetic modeling. Energy Fuel 2014, 28, 6481–6493. [Google Scholar] [CrossRef]
  48. He, L.L.; Yu, D.X.; Lv, W.Z.; Wu, J.Q.; Xu, M.H. CO2 Sequestration by Indirect Carbonation of High-Calcium Coal Fly Ash; Advanced Materials Research, Trans Tech Publications: Geneva, Switzerland, 2013; pp. 2870–2874. [Google Scholar]
  49. He, L.; Yu, D.; Lv, W.; Wu, J.; Xu, M. A novel method for CO2 sequestration via indirect carbonation of coal fly ash. Ind. Eng. Chem. Res. 2013, 52, 15138–15145. [Google Scholar] [CrossRef]
  50. Jo, H.Y.; Kim, J.H.; Lee, Y.J.; Lee, M.; Choh, S.-J. Evaluation of factors affecting mineral carbonation of CO2 using coal fly ash in aqueous solutions under ambient conditions. Chem. Eng. J. 2012, 183, 77–87. [Google Scholar] [CrossRef]
  51. Sun, Y.; Parikh, V.; Zhang, L. Sequestration of carbon dioxide by indirect mineralization using Victorian brown coal fly ash. J. Hazard. Mater. 2012, 209, 458–466. [Google Scholar] [CrossRef]
  52. Ho, H.-J.; Iizuka, A.; Shibata, E.; Ojumu, T. Circular indirect carbonation of coal fly ash for carbon dioxide capture and utilization. J. Environ. Chem. Eng. 2022, 10, 108269. [Google Scholar] [CrossRef]
  53. Lu, L.; Fang, Y.; Huang, Z.; Huang, Y.; Ren, Z.J. Self-sustaining carbon capture and mineralization via electrolytic carbonation of coal fly ash. Chem. Eng. J. 2016, 306, 330–335. [Google Scholar] [CrossRef]
  54. Huang, Y.; Zheng, X.; Wei, Y.; He, Q.; Yan, S.; Ji, L. Protonated amines mediated CO2 mineralization of coal fly ash and polymorph selection of CaCO3. Chem. Eng. J. 2022, 450, 138121. [Google Scholar] [CrossRef]
  55. Zheng, X.; Liu, J.; Wang, Y.; Wang, Y.; Ji, L.; Yan, S. Regenerable glycine induces selective preparation of vaterite CaCO3 by calcium leaching and CO2 mineralization from coal fly ash. Chem. Eng. J. 2023, 459, 141536. [Google Scholar] [CrossRef]
  56. Zheng, X.; Liu, J.; Wei, Y.; Li, K.; Yu, H.; Wang, X.; Ji, L.; Yan, S. Glycine-mediated leaching-mineralization cycle for CO2 sequestration and CaCO3 production from coal fly ash: Dual functions of glycine as a proton donor and receptor. Chem. Eng. J. 2022, 440, 135900. [Google Scholar] [CrossRef]
  57. Ragipani, R.; Sreenivasan, K.; Anex, R.P.; Zhai, H.; Wang, B. Direct Air Capture and Sequestration of CO2 by Accelerated Indirect Aqueous Mineral Carbonation under Ambient Conditions. ACS Sustain. Chem. Eng. 2022, 10, 7852–7861. [Google Scholar] [CrossRef]
  58. Ren, K.; Ma, S.; Feng, Y.; Xu, N.; Bai, S. Study on the composite gravel preparation and the synergistic absorption of CO2 by fly ash of CFB boiler. Fuel 2023, 342, 127843. [Google Scholar] [CrossRef]
  59. Siriruang, C.; Toochinda, P.; Julnipitawong, P.; Tangtermsirikul, S. CO2 capture using fly ash from coal fired power plant and applications of CO2-captured fly ash as a mineral admixture for concrete. J. Environ. Manag. 2016, 170, 70–78. [Google Scholar] [CrossRef]
  60. Chen, T.; Bai, M.; Gao, X. Carbonation curing of cement mortars incorporating carbonated fly ash for performance improvement and CO2 sequestration. J. CO2 Utilizat. 2021, 51, 101633. [Google Scholar] [CrossRef]
  61. Freire, A.L.; Moura-Nickel, C.D.; Scaratti, G.; De Rossi, A.; Araújo, M.H.; Júnior, A.D.N.; Rodrigues, A.E.; Castellón, E.R.; Moreira, R.d.F.P.M. Geopolymers produced with fly ash and rice husk ash applied to CO2 capture. J. Clean. Prod. 2020, 273, 122917. [Google Scholar] [CrossRef]
  62. Yin, T.; Yin, S.; Srivastava, A.; Gadikota, G. Regenerable solvents mediate accelerated low temperature CO2 capture and carbon mineralization of ash and nano-scale calcium carbonate formation. Resour. Conservat. Recycl. 2022, 180, 106209. [Google Scholar] [CrossRef]
  63. Monasterio-Guillot, L.; Alvarez-Lloret, P.; Ibañez-Velasco, A.; Fernandez-Martinez, A.; Ruiz-Agudo, E.; Rodriguez-Navarro, C. CO2 sequestration and simultaneous zeolite production by carbonation of coal fly ash: Impact on the trapping of toxic elements. J. CO2 Utilizat. 2020, 40, 101263. [Google Scholar] [CrossRef]
  64. Yan, F.; Jiang, J.; Liu, N.; Gao, Y.; Meng, Y.; Li, K.; Chen, X. Green synthesis of mesoporous γ-Al2O3 from coal fly ash with simultaneous on-site utilization of CO2. J. Hazard. Mater. 2018, 359, 535–543. [Google Scholar] [CrossRef]
  65. Liu, W.; Su, S.; Xu, K.; Chen, Q.; Xu, J.; Sun, Z.; Wang, Y.; Hu, S.; Wang, X.; Xue, Y. CO2 sequestration by direct gas–solid carbonation of fly ash with steam addition. J. Clean. Prod. 2018, 178, 98–107. [Google Scholar] [CrossRef]
  66. Ćwik, A.; Casanova, I.; Rausis, K.; Koukouzas, N.; Zarębska, K. Carbonation of high-calcium fly ashes and its potential for carbon dioxide removal in coal fired power plants. J. Clean. Prod. 2018, 202, 1026–1034. [Google Scholar] [CrossRef]
  67. Blamey, J.; Manovic, V.; Anthony, E.J.; Dugwell, D.R.; Fennell, P.S. On steam hydration of CaO-based sorbent cycled for CO2 capture. Fuel 2015, 150, 269–277. [Google Scholar] [CrossRef]
  68. González, B.; Liu, W.; Sultan, D.; Dennis, J. The effect of steam on a synthetic Ca-based sorbent for carbon capture. Chem. Eng. J. 2016, 285, 378–383. [Google Scholar] [CrossRef]
  69. Liu, R.; Wang, X.; Gao, S. CO2 capture and mineralization using carbide slag doped fly ash. Greenhouse Gases Sci. Technol. 2020, 10, 103–115. [Google Scholar] [CrossRef]
  70. Chen, H.; Wang, F.; Zhao, C.; Khalili, N. The effect of fly ash on reactivity of calcium based sorbents for CO2 capture. Chem. Eng. J. 2017, 309, 725–737. [Google Scholar] [CrossRef]
  71. Ma, X.; Li, Y.; Yan, X.; Zhang, W.; Zhao, J.; Wang, Z. Preparation of a morph-genetic CaO-based sorbent using paper fibre as a biotemplate for enhanced CO2 capture. Chem. Eng. J. 2019, 361, 235–244. [Google Scholar] [CrossRef]
  72. Li, Y.; Zhao, C.; Ren, Q.; Duan, L.; Chen, H.; Chen, X. Effect of rice husk ash addition on CO2 capture behavior of calcium-based sorbent during calcium looping cycle. Fuel Process. Technol. 2009, 90, 825–834. [Google Scholar] [CrossRef]
  73. Liu, Y.; Yang, X.; Zhao, L.; Lei, F.; Xiao, Y. Effects of steam on CO2 absorption ability of calcium-based sorbent modified by peanut husk ash. Sci. China Technol. Sci. 2017, 60, 953–962. [Google Scholar] [CrossRef]
  74. Yan, F.; Jiang, J.; Li, K.; Tian, S.; Zhao, M.; Chen, X. Performance of Coal Fly Ash Stabilized, CaO-based Sorbents under Different Carbonation–Calcination Conditions. ACS Sustain. Chem. Eng. 2015, 3, 2092–2099. [Google Scholar] [CrossRef]
  75. Sreenivasulu, B.; Sreedhar, I.; Venugopal, A.; Reddy, B.; Raghavan, K. Thermokinetic investigations of high temperature carbon capture using a coal fly ash doped sorbent. Energy Fuel 2017, 31, 6320–6328. [Google Scholar] [CrossRef]
  76. Liu, C.-F.; Yang, C.-H.; Shih, S.-M. CO2 Capture by Fly Ash/Hydrated Lime Sorbents at Low Temperatures. Ind. Eng. Chem. Res. 2022, 61, 4774–4783. [Google Scholar] [CrossRef]
  77. De Aquino, T.F.; Estevam, S.T.; Viola, V.O.; Marques, C.R.; Zancan, F.L.; Vasconcelos, L.B.; Riella, H.G.; Pires, M.J.; Morales-Ospino, R.; Torres, A.E.B. CO2 adsorption capacity of zeolites synthesized from coal fly ashes. Fuel 2020, 276, 118143. [Google Scholar] [CrossRef]
  78. Chen, W.; Song, G.; Lin, Y.; Qiao, J.; Wu, T.; Yi, X.; Kawi, S. A green and efficient strategy for utilizing of coal fly ash to synthesize K-MER zeolite as catalyst for cyanoethylation and adsorbent of CO2. Micropor. Mesopor. Mater. 2021, 326, 111353. [Google Scholar] [CrossRef]
  79. Cheng, S.-Y.; Liu, Y.-Z.; Qi, G.-S. Experimental study of CO2 capture enhanced by coal fly ash-synthesized NH2-MCM-41 coupled with high gravity technology. Chem. Eng. J. 2020, 400, 125946. [Google Scholar] [CrossRef]
  80. Qu, F.; Yan, F.; Shen, X.; Li, C.; Chen, H.; Wang, P.; Zhang, Z. Novel PEI@ CSH adsorbents derived from coal fly ash enabling efficient and in-situ CO2 capture: The anti-urea mechanism of CSH support. J. Clean. Prod. 2022, 378, 134420. [Google Scholar] [CrossRef]
  81. Chen, T.-L.; Chen, Y.-H.; Dai, M.-Y.; Chiang, P.-C. Stabilization-solidification-utilization of MSWI fly ash coupling CO2 mineralization using a high-gravity rotating packed bed. Waste Manag. 2021, 121, 412–421. [Google Scholar] [CrossRef]
  82. Fernandez-Turiel, J.; De Carvalho, W.; Cabañas, M.; Querol, X.; Lopez-Soler, A. Mobility of heavy metals from coal fly ash. Environ. Geol. 1994, 23, 264–270. [Google Scholar] [CrossRef]
  83. Koukouzas, N.; Ketikidis, C.; Itskos, G. Heavy metal characterization of CFB-derived coal fly ash. Fuel Process. Technol 2011, 92, 441–446. [Google Scholar] [CrossRef]
  84. Cardenas-Escudero, C.; Morales-Flórez, V.; Pérez-López, R.; Santos, A.; Esquivias, L. Procedure to use phosphogypsum industrial waste for mineral CO2 sequestration. J. Hazard. Mater. 2011, 196, 431–435. [Google Scholar] [CrossRef]
  85. Chen, J.; Fu, C.; Mao, T.; Shen, Y.; Li, M.; Lin, X.; Li, X.; Yan, J. Study on the accelerated carbonation of MSWI fly ash under ultrasonic excitation: CO2 capture, heavy metals solidification, mechanism and geochemical modelling. Chem. Eng. J. 2022, 450, 138418. [Google Scholar] [CrossRef]
Figure 1. Flow chart of direct and indirect CO2 mineralization using COFA [13]. Copyright 2012, with permission from Elsevier.
Figure 1. Flow chart of direct and indirect CO2 mineralization using COFA [13]. Copyright 2012, with permission from Elsevier.
Energies 16 06241 g001
Figure 2. Carbonation reaction mechanism of COFA directly mineralizes CO2 [20]. Copyright 2023, with permission from Elsevier.
Figure 2. Carbonation reaction mechanism of COFA directly mineralizes CO2 [20]. Copyright 2023, with permission from Elsevier.
Energies 16 06241 g002
Figure 3. (a) Sketch map of measuring the CO2 diffusion distance within block CaO, morphology images of (b,c) non-supercritical CO2, (d,e) supercritical CO2 [25]. Copyright 2022, with permission from Elsevier.
Figure 3. (a) Sketch map of measuring the CO2 diffusion distance within block CaO, morphology images of (b,c) non-supercritical CO2, (d,e) supercritical CO2 [25]. Copyright 2022, with permission from Elsevier.
Energies 16 06241 g003
Figure 4. Procedure for the determination of carbonation depth [29]. Copyright 2015, with permission from Elsevier.
Figure 4. Procedure for the determination of carbonation depth [29]. Copyright 2015, with permission from Elsevier.
Energies 16 06241 g004
Figure 5. Schematic diagram of integrated CO2 sequestration, circular fluidized bed fly ash stabilization, and by-product application via high-gravity carbonation process [32]. Copyright 2016, with permission from the American Chemical Society.
Figure 5. Schematic diagram of integrated CO2 sequestration, circular fluidized bed fly ash stabilization, and by-product application via high-gravity carbonation process [32]. Copyright 2016, with permission from the American Chemical Society.
Energies 16 06241 g005
Figure 6. Mechanism of high-efficiency CO2 mineralization using COFA based on amine-looping [44]. Copyright 2022, with permission from Elsevier.
Figure 6. Mechanism of high-efficiency CO2 mineralization using COFA based on amine-looping [44]. Copyright 2022, with permission from Elsevier.
Energies 16 06241 g006
Figure 7. Process flow sheet of the integrated process combining CO2 absorption with mineralization [45]. Copyright 2018, with permission from Elsevier.
Figure 7. Process flow sheet of the integrated process combining CO2 absorption with mineralization [45]. Copyright 2018, with permission from Elsevier.
Energies 16 06241 g007
Figure 8. Impact of carbonation temperature on CO2 uptake at 10 bar [51]. Copyright 2012, with permission from Elsevier.
Figure 8. Impact of carbonation temperature on CO2 uptake at 10 bar [51]. Copyright 2012, with permission from Elsevier.
Energies 16 06241 g008
Figure 9. The integrated system of circular indirect carbonation of coal fly ashes [52]. Copyright 2022, with permission from Elsevier.
Figure 9. The integrated system of circular indirect carbonation of coal fly ashes [52]. Copyright 2022, with permission from Elsevier.
Energies 16 06241 g009
Figure 10. Schematic diagram of the electrolytic carbonation of COFA for in situ CO2 sequestrate and mineral recovery [53]. Copyright 2016, with permission from Elsevier.
Figure 10. Schematic diagram of the electrolytic carbonation of COFA for in situ CO2 sequestrate and mineral recovery [53]. Copyright 2016, with permission from Elsevier.
Energies 16 06241 g010
Figure 11. Surface morphology images of CaCO3 induced by 11 types of selected amines [54]. Copyright 2022, with permission from Elsevier.
Figure 11. Surface morphology images of CaCO3 induced by 11 types of selected amines [54]. Copyright 2022, with permission from Elsevier.
Energies 16 06241 g011
Figure 12. Mechanism of cyclic glycine-mediated leaching-mineralization approach using glycine and COFA [56]. Copyright 2022, with permission from Elsevier.
Figure 12. Mechanism of cyclic glycine-mediated leaching-mineralization approach using glycine and COFA [56]. Copyright 2022, with permission from Elsevier.
Energies 16 06241 g012
Figure 13. Schematic diagram of an integrated system coupling mineral carbonation with direct air capture under atmospheric conditions [57]. Copyright 2022, with permission from the American Chemical Society.
Figure 13. Schematic diagram of an integrated system coupling mineral carbonation with direct air capture under atmospheric conditions [57]. Copyright 2022, with permission from the American Chemical Society.
Energies 16 06241 g013
Figure 14. (a) Photo of composite gravels, (b) compressive strength of composite gravel with different curing durations, (c) CO2 absorption capacity of composite gravel with different gypsum/COFA ratios [58]. Copyright 2023, with permission from Elsevier.
Figure 14. (a) Photo of composite gravels, (b) compressive strength of composite gravel with different curing durations, (c) CO2 absorption capacity of composite gravel with different gypsum/COFA ratios [58]. Copyright 2023, with permission from Elsevier.
Energies 16 06241 g014
Figure 15. Schematic diagram of an integrated approach combining CO2 capture with mineralization adopting sodium glycinate as a regenerable solvent and CTAB surfactant to control the size of CaCO3 particles [62]. Copyright 2022, with permission from Elsevier.
Figure 15. Schematic diagram of an integrated approach combining CO2 capture with mineralization adopting sodium glycinate as a regenerable solvent and CTAB surfactant to control the size of CaCO3 particles [62]. Copyright 2022, with permission from Elsevier.
Energies 16 06241 g015
Figure 16. Mesoporous γ-Al2O3 synthesized from COFA combined with simultaneous on-site CO2 utilization [64]. Copyright 2018, with permission from Elsevier.
Figure 16. Mesoporous γ-Al2O3 synthesized from COFA combined with simultaneous on-site CO2 utilization [64]. Copyright 2018, with permission from Elsevier.
Energies 16 06241 g016
Figure 17. CO2 sequestration capacities of dry COFA [40]. Copyright 2017, with permission from Elsevier.
Figure 17. CO2 sequestration capacities of dry COFA [40]. Copyright 2017, with permission from Elsevier.
Energies 16 06241 g017
Figure 18. Diagram of the pilot-scale CEB bed reactor system for carbonation of fly ash [66]. Copyright 2018, with permission from Elsevier.
Figure 18. Diagram of the pilot-scale CEB bed reactor system for carbonation of fly ash [66]. Copyright 2018, with permission from Elsevier.
Energies 16 06241 g018
Figure 19. CaO sorbents doped with nano silica derived from COFA used for CO2 capture [17]. Copyright 2017, with permission from the American Chemical Society.
Figure 19. CaO sorbents doped with nano silica derived from COFA used for CO2 capture [17]. Copyright 2017, with permission from the American Chemical Society.
Energies 16 06241 g019
Figure 20. Micrographs of the zeolites synthesized from COFA (a) synthetic SZX, (b) PR1, (c) NaAFA and (d) HP1 zeolites. Modification from refs. [14,15,77,78].
Figure 20. Micrographs of the zeolites synthesized from COFA (a) synthetic SZX, (b) PR1, (c) NaAFA and (d) HP1 zeolites. Modification from refs. [14,15,77,78].
Energies 16 06241 g020
Figure 21. Mechanisms of CO2 absorption improvement of nanofluids in high gravity rotating packed bed [79]. Copyright 2020, with permission from Elsevier.
Figure 21. Mechanisms of CO2 absorption improvement of nanofluids in high gravity rotating packed bed [79]. Copyright 2020, with permission from Elsevier.
Energies 16 06241 g021
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

Jiang, L.; Cheng, L.; Zhang, Y.; Liu, G.; Sun, J. A Review on CO2 Sequestration via Mineralization of Coal Fly Ash. Energies 2023, 16, 6241. https://0-doi-org.brum.beds.ac.uk/10.3390/en16176241

AMA Style

Jiang L, Cheng L, Zhang Y, Liu G, Sun J. A Review on CO2 Sequestration via Mineralization of Coal Fly Ash. Energies. 2023; 16(17):6241. https://0-doi-org.brum.beds.ac.uk/10.3390/en16176241

Chicago/Turabian Style

Jiang, Long, Liang Cheng, Yuxuan Zhang, Gaojun Liu, and Jian Sun. 2023. "A Review on CO2 Sequestration via Mineralization of Coal Fly Ash" Energies 16, no. 17: 6241. https://0-doi-org.brum.beds.ac.uk/10.3390/en16176241

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