Next Article in Journal
Rooftop Solar PV Policy Assessment of Global Best Practices and Lessons Learned for the Kingdom of Saudi Arabia
Previous Article in Journal
Digital Model of Deflection of a Cable-Stayed Bridge Driven by Deep Learning and Big Data Optimized via PCA-LGBM
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress in the Preparation of Calcium Carbonate by Indirect Mineralization of Industrial By-Product Gypsum

School of Petrochemical Technology, Lanzhou University of Technology, Lanzhou 730050, China
*
Authors to whom correspondence should be addressed.
Sustainability 2023, 15(12), 9629; https://0-doi-org.brum.beds.ac.uk/10.3390/su15129629
Submission received: 13 April 2023 / Revised: 25 May 2023 / Accepted: 14 June 2023 / Published: 15 June 2023

Abstract

:
To avoid the long-term pollution of land and water by industrial gypsum by-products, the exploitation of this resource has become a priority. The indirect synthesis of calcium carbonate from the industrial by-product gypsum has received substantial attention as a viable method for resource utilization. Currently, the primary problems in the indirect manufacture of calcium carbonate from the industrial by-product gypsum are additive recycling and process simplification. This paper describes the present state of development and compares various indirect mineralization systems. The factors affecting leaching and mineralization in the indirect mineralization of CO2 from by-product gypsum and the management of CaCO3 crystallinity are discussed, and the current additive regeneration cycle is summarized. The applications of other technologies in the indirect mineralization of by-product gypsum are also summarized, as are the obstacles, and required future work. This review provides guidelines for the laboratory indirect mineralization of by-product gypsum as well as practical applications.

1. Introduction

Owing to its low cost, abundance, and excellent biocompatibility, calcium carbonate (CaCO3) is a frequently used filler in the paper, plastic, rubber, and food sectors [1,2,3]. Carbonation [4], complicated decomposition [5], and pyrolysis [6] are the most common techniques used to produce calcium carbonate. Although well developed, the conventional production methods have drawbacks such as high energy consumption and the depletion of natural ores [2,7,8]. To address these drawbacks, a novel method for producing calcium carbonate using calcium-containing industrial solid waste as a raw material rather than natural minerals has been developed [9,10]. Reacting CO2 from flue gas with calcium from the industrial by-product gypsum (CaSO4) produces calcium carbonate more economically than the traditional approach does by eliminating the additional stages of mining, grinding, and separation of natural minerals [11]. Additionally, the industrial by-product gypsum can be obtained from a variety of sources located close to the source of CO2 emissions, such as thermal power plants [12]. In addition to being an innovative application of gypsum, this method also sequesters CO2.
Three significant industrial gypsum by-products are flue-gas desulfurization gypsum (FGDG) [13], which is produced from coal combustion after the wet desulfurization of flue gas; phosphogypsum (PG) [14], which originates from phosphate fertilizer production; and red gypsum (RG) [15], a by-product of the neutralization of spent sulfuric acid in titanium dioxide production. In 2020, China produced approximately 200 Mt of industrial gypsum by-products, 90% of which was FGDG, phosphogypsum, and titanium gypsum [16]. Similarly, in the EU and the US, phosphogypsum from fertilizer production is a gigantic deal [17]. The major components and chemical properties of by-product gypsum are identical to those of natural gypsum, and it has been extensively researched for usage as construction materials [18], extraction of chemical products [19,20,21], calcium sulfate whiskers manufacture [22], soil remediation [23], and so on. Despite the huge production of gypsum by-products, their utilization rate in China is very low, approximately 72% for desulfurization gypsum [24], 40% for phosphogypsum [14] and less than 10% for red gypsum [25]. Thus, high value and large-scale applications for gypsum byproducts must be urgently developed.
According to different technical procedures, mineralization of by-product gypsum may be classified as direct mineralization or indirect mineralization. The direct aqueous phase mineralization method involves the simultaneous reaction of by-product gypsum, alkali (e.g., ammonia(NH3·H2O) and sodium hydroxide (NaOH)) and CO2 gas [26,27,28]. The indirect aqueous phase mineralization method extracts the reactive component (Ca2+) from by-product gypsum using acids, bases, and other extracting agents, and then carbonates with CO2 in an aqueous solution [29]. Compared with direct mineralization, the industrial application of indirect mineralization using by-product gypsum as the raw material is slow. In China, a pilot plant for direct mineralization with phosphogypsum and CO2 as raw materials in natural gas, which was completed by the Sinopec Group and Sichuan University, was successfully operated in the Puguang Natural Gas Purification Plant [30]. Sinochem Chongqing Fuling Chemical Co., Ltd., cooperated with the Chinese Academy of Sciences to build a demonstration plant with an annual output of 100,000 tons of high-concentration phosphogypsum mineralization capacity to achieve a stable operation [31]. Therefore, the industrial application of indirect mineralization of by-product gypsum will also be the main research direction in the future.

2. Chemical Composition of Industrial By-Product Gypsum and Principles of Indirect Mineralization

The main component of industrial by-product gypsum is CaSO4·2H2O. However, different impurities arise from different production processes and sources [32,33]. Figure 1 shows the scanning electron microscopy (SEM) images of the gypsum by-products of the three industrial processes. As shown in Table 1, the main components of flue gas desulfurization gypsum are SO3, CaO, and bound water [34].
The main components of phosphogypsum and red gypsum are similar to those of FGDG, but they contain more impurities. In addition to other minor impurities, phosphogypsum can contain silica, free H3PO4 [38], and phosphate, whereas red gypsum contains iron oxide (Fe2O3) and titanium oxide (TiO2) [39]. The sulfur contents of FGDG, phosphogypsum, and red gypsum range from 12.80 to 18.60 wt.%, 11.99 to 18.40 wt.%, and 12.64 to 15.42 wt.%, respectively, with similar ranges. However, the different calcium and impurity contents indicate differences among the three types of waste gypsum.
Table 1. Chemical composition of industrial by-product gypsum.
Table 1. Chemical composition of industrial by-product gypsum.
FGDG [40]FGDG [41]FGDG [42]PG [43]PG [36]PG [44]RG [45]RG [29]RG [46]
OriginSpainTurkeyChinaLithuaniaKazakhstanSpainMalaysiaMalaysiaItaly
wt.%(CaO)42.2032.5032.2538.6028.54 32.00 32.2033.1229.30
wt.% (SO3)54.3046.5142.58 53.4829.97 46.00 31.6029.9338.45
wt.% (SiO2)<0.1-0.78 0.3715.402.52 1.90-2.31
wt.% (Na2O)----4.440.01--0.74
wt.% (Al2O3)0.600.560.120.131.13 0.20 0.70-1.25
wt.% (Fe2O3)0.300.430.320.030.85 -28.9929.233.97
wt.% (MgO)0.17-0.520.040.66 ---2.46
wt.% (P2O5)---0.822.52 0.65 ---
wt.% (F) ---0.147.90 ---0.03
wt.% (TiO2) -0.02--0.16 -5.646.211.29
wt.% (MnO)-0.03--0.15-0.410.58-
wt.% (Eu2O3)------0.260.34-
wt.% (ZnO)------0.240.19-
wt.% (V2O5)------0.220.170.18
wt.%
(Crystalline water)
23.0019.7018.346.40-18.40--19.05
Table 1 shows that by-product gypsum contains a large amount of calcium (22.38–34.13 wt.%). As the solubility product of calcium carbonate (Ksp = 3.36 × 10−9) [47] is much lower than that of calcium sulfate (Ksp = 3.14 × 10−5) [48], by-product gypsum has the potential to be mineralized to calcium carbonate. As the acidity of CO2 dissolved in water is not sufficient to dissolve minerals [49,50], various physical and chemical schemes have been proposed to facilitate the dissolution of calcium from by-product gypsum [51,52]. Calcium carbonate was prepared from by-product gypsum via indirect mineralization using the following steps: First, Ca2+ was extracted from the gypsum by-product using leaching agents (e.g., ammonium salts (NH4Cl) and acid and bases (such as H2SO4 and NaOH). Next, CO2 was passed through the calcium-rich solution to form calcium carbonate and complete the CO2 sequestration. Finally, the leaching agent was regenerated and recycled [53,54,55]. The technical process is shown in Figure 2, corresponding to reaction Equations (1)–(3).
CaSO4·2H2O(s) + Leaching agents ↔ Ca2+ + SO42− + 2H2O
2OH + CO2(g) → CO32− + H2O
Ca2+ + CO32− → CaCO3(s)

3. Main Technical Routes and Research Progress of Indirect Mineralization of By-Product Gypsum

3.1. Main Technical Routes for Indirect Mineralization of By-Product Gypsum

The preparation of calcium carbonate via the indirect mineralization of industrial byproduct gypsum can be divided into four processes: alkali leaching, pH swing, salt leaching, and complexation.

3.1.1. Alkali Leaching

Blencoe et al. [56] first proposed the use of NaOH to leach calcium and magnesium from silicate minerals to prepare carbonates via mineralization and CO2 sequestration. Because calcium hydroxide is less alkali than sodium hydroxide or ammonia, the calcium in the industrial by-product gypsum can be converted to Ca(OH)2 by the action of the alkali. According to the principle of making a weak alkali from a strong alkali, the transfer of calcium from calcium hydroxide to calcium carbonate is then achieved via the passage of CO2. Corresponding to reaction Equations (4)–(6), the process of preparing calcium carbonate via the indirect mineralization of phosphogypsum with NaOH as a leaching agent is shown in Figure 3.
CaSO4·2H2O(s) + 2OH(aq) ↔ Ca(OH)2(s) + SO42−(aq) + 2H2O(l)
CaSO4·2H2O(s) + 2NH3·H2O ↔ Ca(OH)2(s) + 2NH4(aq) + SO42−(aq) + 2H2O(l)
Ca(OH)2(s) + CO2(g) → CaCO3(s) + H2O(l)
The type of base plays an important role in leaching, mineralization efficiency, and calcium carbonate purity. Pérez-Moreno et al. [57] and Cárdenas-Escudero et al. [51] found that the conversion of red gypsum to calcium carbonate was 92% while leaching with sodium hydroxide was 64% while leaching with ammonia. However, the impurities and heavy metals in red gypsum are also transferred during the conversion to Ca(OH)2, which ultimately results in a lower purity of calcium carbonate [57,58,59]. Romero-Hermida et al. [58] prepared calcium carbonate by dissolving phosphogypsum in corrosive aqueous waste from aluminum anodizing based on the principle of alkali leaching. Corresponding to reaction Equations (7) and (8), a mineralization efficiency of 80% was obtained.
3Ca(OH)2 + 12OH(aq) + 2Al3+(aq) ↔ 3SO42−(aq) + Ca3Al2(OH)12(s) + 6H2O(l)
Ca3Al2(OH)12(s) + 3CO2(g) ↔ 3CaCO3(s) + 2Al(OH)3(s) + 3H2O(l)
In addition, the type of alkali used has an important influence on the calcium carbonate crystal form. Pérez-Moreno et al. [57] found that calcium carbonate produced after alkali leaching with NaOH contained 70–80% calcite, and its particles were poorly developed and rectangular or diamond shaped. However, the main crystalline phases of calcium carbonate after ammonia leaching were gypsum (50–55%) and calcite (25–30%).
Based on the principle of alkali leaching, Altiner et al. [60] investigated the effect of reactor design on the mineralization time and change in the calcium carbonate particle size. The reaction time required to produce calcium carbonate particles in a Venturi reactor was half of that without the Venturi tube, and all the calcium carbonate particles produced were calcite. However, the crystal structure was affected by the experimental conditions.
As shown in Table 2, the alkali leaching of gypsum byproducts does not use additives, and the final calcium carbonate crystal formed is mainly calcite. However, the mineralization efficiency and calcium carbonate purity differ because of the difference in gypsum by-products. In addition, the indirect mineralization of the gypsum byproduct and other wastes (such as caustic waste liquid [58], bauxite slag [61], waste brine [62,63], and seawater [64]) will play a significant role in the comprehensive utilization of waste resources in the future. However, alkali leaching still requires a large amount of alkali liquor, which is corrosive to equipment, and the remaining alkali and Na2SO4 in the filtrate after mineralization are difficult to separate. Therefore, achieving efficient circulation is difficult.

3.1.2. pH Swing

Extracting calcium and magnesium ions from raw minerals or alkali solid waste is easier in acidic environments than in alkali environments; however, high-purity carbonate precipitates can be prepared in alkali environments [65,66]. Changing the pH from low to high or from high to low during leaching–mineralization is called the pH swing method. Alissa et al. [67] first mineralized and sequestered CO2 from serpentine using the pH swing method to obtain high-purity magnesium carbonate. Preparing calcium carbonate via the pH swing indirect mineralization of by-product gypsum includes dissolving calcium or iron at low pH to separate insoluble impurities and then adding NH3·H2O to increase the pH to remove the main impurities (iron or silicon). Finally, carbonation of the high-purity Ca-rich solution with CO2 is performed at pH 10. Figure 4 shows a schematic of the indirect mineralization of red gypsum using the pH swing method.
Different leaching agents have different leaching efficiencies for gypsum by-products during the pH swing process. Amin et al. [69] compared the effects of acid and alkali leaching agents on red gypsum. Acidic leaching agents (H2SO4, HCl, and HNO3) extracted calcium and iron more efficiently than alkali leaching agents (NaOH, KOH, and NH3·H2O) did. Additionally, HCl extracted calcium more efficiently than HNO3 and H2SO4 did. Thus, calcium carbonate from red gypsum mineralization is generally obtained via the swing process of adding H2SO4 for leaching and then adding an alkali to adjust the pH [10,12,42].
The process conditions also significantly affect the reaction kinetics and crystal form of the calcium carbonate. Amin et al. [70] observed that an equilibrium always exists between Ca2+ and CO2 during mineralization. At a specific calcium concentration, increasing the CO2 pressure improves the mineralization efficiency. They also found that the mineralization efficiency was much higher while using ammonium bicarbonate as a CO32− source rather than CO2 [71]. At the same time, it is confirmed that the mixture of vaterite, calcite and aragonite can be formed below 80 °C, while vaterite alone can be formed only above 170 °C [71]. To promote the dissolution of CO2 in the mineralization stage, Omeid et al. [29] added monoethanolamine (MEA) as a CO2 absorbent, which yielded a mineralization efficiency of 98.9% for calcite calcium carbonate. A positive correlation also existed among the mineralization efficiency of calcite, MEA consumption, and CO2 dissolution. Ding et al. [72] found that, in the pH swing process of phosphogypsum, vaterite can be prepared at a lower temperature (20–40 °C) when ammonia water is added during mineralization. As the temperature increases, some of the vaterite is transformed to aragonite and calcite, decreasing the purity of vaterite (Figure 5).
Amin et al. [65,69,70,71,73,74] used the reaction model PHREEQC-2.18 program to simulate the mineralization and precipitation of various solid phases during the indirect mineralization of red gypsum via the pH swing method, so as to predict CO2 absorption. The results showed that calcium ions were rapidly released from CaO at pH 12.5. In addition, the simulation results directly showed the saturation indices of CaO and CaCO3, indicating that their dissolution and precipitation depend on the saturation state and pH value.
Ding et al. [72] removed the main impurity in phosphogypsum, SiO2, using the pH swing process of alkali leaching, acid washing, and mineralization in an alkali environment. The calcium leaching rate reached 99.6%, and the mineralization efficiency reached 98.57%. The relevant reactions are shown in Equations (9)–(11).
Ca(OH)2(s) → Ca2+(aq) + 2OH(aq)
Ca2+(aq)+ 2NH3·H2O → Ca(OH)2(s) + 2NH4+(aq)
Ca(OH)2(s) + CO2(g) → CaCO3(s) + H2O(l)
As shown in Table 2, in the pH swing process of by-product gypsum, the conditions without additives in the mineralization stage are more stringent than those with additives. Therefore, considering the energy consumption and cost, a more efficient additive selection or adjustment to produce higher value and purer calcium carbonate will be a future development trend. Although the pH swing process for leaching by-product gypsum improves leaching efficiency and removes impurities [15], it has obvious disadvantages. First, the process is complex and requires large amounts of acids and alkalis. Second, recycling the leaching agent is difficult, which is expensive and harmful to the environment.

3.1.3. Salt Leaching

The addition of a strongly soluble electrolyte to increase the solubility of insoluble substances is called the salt effect. Salt leaching applies this principle to the indirect mineralization of gypsum by-products to calcium carbonate. The gypsum byproduct extracted via the salt effect is mainly divided into ammonium salt and other salts according to the salt type. Chen et al. [54] proposed leaching calcium from phosphogypsum using sodium chloride followed by mineralization in an ammonia water environment to prepare calcium carbonate. For reaction equations such as 13 and 14, the results show that under the optimized conditions (3 mol/L, 50: 1 mL/g, 30 °C, 60 min), the leaching rate of Ca2+ is 49.42% and the carbonation rate is 96.31%. Ding et al. [75] used ammonium acetate to separate Ca2+ from PG and concentrate impurity ions. For reaction Equations (14)–(18), the leaching and mineralization efficiencies of calcium were more than 98%. The phosphogypsum was leached with NH4Cl and mineralized in ammonia water. The results showed that the optimum dissolved amount of CaSO4·2H2O was 18.7 g/L and the carbonation rate was 98.22% [76].
PG(s) → Ca2+(aq) + SO42−(aq) + SiO2(s)
Ca2+(aq) + NH3·H2O + CO2(g) → CaCO3(s) + NH4+(aq) + H2O(l)
CaSO4(s) +2CH2COO(aq) → Ca(CH2COO)2+ SO42−(aq)
Ca(CH2COO)2 +2NH3·H2O → Ca(OH)2(s) + 2CH2COO(aq) + 2NH4+(aq)
Ca(OH)2(s) + CO2(g) → CaCO3(s) + H2O(l)
2 NH3·H2O + CO2(g) → 2NH4+(aq) + CO32−(aq) + H2O(l)
Ca2+(aq) + CO32−(aq) → CaCO3(s)
The regenerative stability also differed after leaching and mineralization with different salts. When NaCl was used as the leaching agent, it could be recycled four times, and the corresponding reaction efficiency could exceed 60% [54]. When NH4Cl was used as the leaching agent, the mineralized filtrate was recycled up to nine times, as shown in Figure 6a, and the reaction efficiency was maintained above 50%. The crystal phase of the obtained product was calcite, and no other phases of CaCO3 were found (Figure 6b) [76,77].
In salt leaching, the experimental conditions also have an important influence on the mineralization efficiency and final calcium carbonate morphology. Chen et al. [54] found that during the mineralization of phosphogypsum in a NaCl–NH3·H2O system, increasing the recovery time of the mineralized filtrate caused the morphology of the calcium carbonate to change gradually from calcite to vaterite. Ding et al. [75] first produced pure vaterite calcium carbonate at 20 °C in an ammonia water environment. With increasing temperature, vaterite was converted into aragonite without any crystal-form regulator [75]. This confirms that increasing the amount of ammonia is beneficial for vaterite formation [54,75,76]. Additionally, the mineralization efficiency decreased with increasing temperature, and the effects of ammonia addition, CO2 gas velocity, and reaction time on the mineralization efficiency increased to a plateau.
To further study the mineralization reaction mechanism of phosphogypsum in the CH3COONH4–NH3·H2O system, Ding et al. [75] calculated the thermodynamic parameters under standard reaction conditions. The results show that the three-step reaction mechanism of phosphogypsum mineralization in this system is as follows: (a) leaching Ca2+ from the original phosphogypsum, (b) conversion of calcium acetate into calcium hydroxide via ammonia water, and (c) conversion of calcium hydroxide into calcium carbonate via the dissolution of gaseous CO2 (g) in water (Figure 7).
Although as mentioned above, leaching agents such as NaCl and NH4Cl can be regenerated and recycled. Figure 6 shows that the leaching efficiency of the mineralized filtrate gradually decreases as the number of cycles increases. Therefore, the development of additives with better regenerative stabilities is an important objective for future research on the indirect mineralization of byproduct gypsum to calcium carbonate.

3.1.4. Complexation

The application and kinetic analysis of organic acids containing carboxyl or hydroxyl groups that can promote mineral dissolution via complexation has been extensively researched. Fredd et al. [78] found that chelators such as cyclohexanediamine tetraacetic acid (CDTA), diethyltriamine pentaacetic acid (DTPA), and ethylenediamine tetraacetic acid (EDTA) significantly increased the dissolution of calcite. The chelating agent promotes the dissolution of the mineral phase via a ligand-exchange reaction, yielding a highly soluble metal complex [79].
Owing to their different complexation constants for calcium ions, different organic acids have different efficiencies for leaching and final mineralization of the same raw material [80], and they produce different purities and calcium carbonate morphologies [81]. Yang et al. [53] used sodium gluconate as a phase-transfer agent to indirectly mineralize phosphogypsum and prepare calcium carbonate microspheres. The presence of sodium gluconate inhibited the nucleation and growth of calcite but promoted the formation of vaterite. The interaction between sodium gluconate and Ca2+ plays a key role in the formation of monodisperse vaterite CaCO3. These are summarized in reactions 19 and 20. Yuan et al. [82] used the same “phase transfer–precipitation” route to mineralize gypsum scales attached to evaporator walls to prepare calcium carbonate micron rods.
CaSO4·2H2O(s) + 2C6H11O7(aq) → Ca(C6H11O7)2 + SO42−(aq) + H2O(l)
Ca(C6H11O7)2 + CO2(g) + 2OH(aq) → CaCO3(s) + 2C6H11O7(aq) + H2O(l)
Both the carboxyl and amino groups of amino acids form complexes with calcium ions. However, researchers have long focused on the regulation of amino acids in the crystal form of calcium carbonate during mineralization [83,84]. In recent years, amino acids have attracted considerable attention as new additives for leaching–mineralization cycles. Zheng et al. [85,86,87] found that amino acids act as not only leaching agents via complexation effects in the leaching stage of indirect mineralization but also CO2 absorbers and calcium carbonate crystal inducers in the mineralization stage. The related reaction Equations are (21)–(23).
CaO(s) + H2O(l) → Ca(OH)2
Ca(OH)2 + 2NH2CH2COO(aq) +2H+(aq) ↔ Ca(NH2CH2COO)2 + 2H2O(l)
Ca(NH2CH2COO)2 + CO2(g) ↔ CaCO3(s) + 2NH2CH2COO(aq) +2H+(aq)
Because the main component of gypsum is CaSO4·2H2O, the filtrate after mineralization is enriched with SO42−. Thus, if sulfate and the leaching agent are not separated before recycling, the dissolution of CaSO4·2H2O [30] will be inhibited by the common-ion effect. This is also the main reason why the leaching efficiency gradually decreases as the number of cycles increases in salt leaching.
Gong et al. [55] leached and mineralized CaSO4·2H2O by exploiting the solubility difference of aspartic acid at different pH values. The results showed that the solubility of CaSO4·2H2O was 16.3 g L−1 in 7 wt.% ammonia and a liquid-to-solid ratio of 50. Owing to the absorption of CO2 by aspartic acid during mineralization, the induction time of calcium carbonate crystallization was significantly increased (as shown in Figure 8b), and the obtained calcium carbonate was a homogeneous vaterite type (as shown in Figure 8c), which confirmed that a relatively high content of amino acids could aid in the formation of the vaterite phase. The process route is illustrated in Figure 8a.
The regeneration of the leaching agent is also different when the industrial by-product gypsum is indirectly mineralized via complexation. Gong et al. [55] combined the dissolution characteristics of amino acids with the leaching–mineralization of desulfurized gypsum. Because amino acids are sparingly soluble at their isoelectric points, aspartic acid can be precipitated and then recycled by adjusting the pH to the isoelectric point after mineralization. After 10 cycles, the recovery efficiency was maintained at 80%, while the dissolved CaSO4·2H2O and total carbonation efficiency were 16.3 ± 0.4 g L−1 and 46.5% ± 1.9%, respectively (as shown in Figure 9a), and the spherical vaterite particles with diameters of 2–5 mm were recovered after each stage (as shown in Figure 9b).
The complexation of the gypsum by-product will be the preferred choice for indirect mineralization in the future because it uses significantly less acid and alkali, is a multifunctional additive (leaching agent, CO2 absorbent, crystal form regulator), and has a high efficiency of regeneration and circulation. However, the leaching kinetics of complexation remain unclear, and the complexation via aspartic acid requires further study.
Table 2. Technical parameters and calcium carbonate morphology of indirect mineralization process of by-product gypsum.
Table 2. Technical parameters and calcium carbonate morphology of indirect mineralization process of by-product gypsum.
Craft
Classification
IngredientsLeachingLeaching Rate (%)MineralizationMineralization Efficiency (%)Calcium CarbonatePurity (%)Ref.
Leaching AgentConditionsAdditivesConditionsCrystallineAppearance
Alkali leachingPGNaOHRT; ATP; PG/H2O mass ratio = 4; OH/Ca2+ molar ratio = 2;3 h97-RT; ATP 20 cm3/s; 20; 15 min; Aging 24 h95CalciteFine-grained>90[88]
PGNaOHOH/Ca molar ratio =2; RT; 3 h--RT; ATP 20 cm3/s; 20; 15 min80Calcite-90[51]
FGDGNaOHS/L = 1:13, 300 rpm; 20 °C; 1 h--RT; 1 L/min; 32.32 m/s Venturi reactor-CalciteRhombus99.63[60]
PGNaOH3 h, 25.0 mL Liquid waste; 12.5 g PG, pH 12.090-RT; ATP 20 cm3/s; 20; 90 min; pH = 6.780Calcite--[58]
pH swing RGH2SO4
NH3·H2O
25 °C; ATP; 600 rpm; 2.5 h; NH3·H2O regulating pH5200 ppm5 wt% MEART; ATP 20 vol% CO2; 1.4 M; 4 h; 600 rpm98.9CalciteRhombus-[29]
RGH2SO4
NH3·H2O
25 °C; ATP; 600 rpm; 2.5 h; NH3·H2O regulating pH--25–150 °C; 1–30% Split pressure CO2; 400 rpm; 3 h98.8CalciteRhombus-[68]
RGH2SO4
NH3·H2O
25 °C; 10 bar; 1000 rpm; 0.5 h; NH3·H2O regulating pH100-25 °C 8 bar; 50 mL; 400 rpm; 30 min100CalciteCube98[70]
RGH2SO4
NH3·H2O
2 M H2SO4; 70 °C; 1000 rpm; 1 h; NH3·H2O regulating pH100NH4HCO330 min; 75 °C; ATP98CalciteRhombus92.57[71]
RGH2SO4
NH3·H2O
70 °C; 1.5 M H2SO4; 2 h; 1 bar; NH3·H2O regulating pH95.8-RT; ATP; 8 mL/min CO2; 50 min93Calcite--[59]
PGNaOH, HCl2 mol/L NaOH; HCl 3 mol/L; 60 °C;10 m L/g; 1 h99.6NH3·H2O30 °C; ATP; 20 mL; 0.1 L/min; 1h98.57SpheruliteSphere99.4[72]
Salting leachingPGNaCl30 °C, ATP; L/S = 50; 60 min49.42NH3·H2O30 °C; ATP; CO2: 80 mL/min; 60 min96.31CalciteRhombus-[54]
PGNH4Cl2 mol/L NH4Cl; 10:1 mL/g; 60 min; 60 °C18.7 g/LNH3·H2O30 °C; ATP; CO2: 50 mL/min; 1h98.22CalciteRhombus99.8[76]
PGCH3COONH46 mol/L; 80 °C; 10 mL/g; 60 min98.1NH3·H2O30 °C; ATP; 100 mL/min; 90 min98.32SpheruliteSphere99.6[75]
FGDGNH4Cl4 mol/L NH4Cl; 80:1 mL/g; 60 min; 50 °C95.6NH3·H2O40 °C; ATP; CO2:80 mL/min; 1 h98.54CalciteRhombus-[77]
ComplexationFGDGAsp30 °C; Asp/BG = 2.5; ATP; L/S = 30; 60 min48.9-30 °C; ATP; 0.5 L/min; 50 min46.5SpheruliteSphere-[55]
PGSodium gluconateSG (nSG/nPGCa2+ = 1:2.5); 100 mL H2O; 9.88 g PG; 20 min84sodium triphosphate15 °C; ATP; (nNaOH/nadditive/nCa) = 2:0.3:1; 300 rpm; 0.4 L/min; Aging 45 min96SpheruliteSphere96.87[53]
Note: RT: room temperature; ATP: atmospheric.

3.2. Application of Other Technologies in the Indirect Mineralization of Industrial By-Product Gypsum

Other technologies have also been applied to improve the indirect mineralization of by-product gypsum. If the sulfate solution produced after mineralization is not used, it is equivalent to producing secondary waste. Recently, the application of bipolar membrane electrodialysis (BMED) to waste salts has attracted considerable attention [89,90]. Ho et al. [91] regenerated a mineralized NaNO3 solution using BMED and obtained NaOH and HNO3 solutions for recycling, which further improved the process economy (Figure 10). Many studies have shown that combining BMED with indirect mineralization is promising [92,93].
To make the reaction more efficient or to control the crystal form of the calcium carbonate produced, researchers have used bubble, Venturi, and membrane electrolysis reactors to indirectly mineralize gypsum by-products. Altiner et al. [60] applied a microbubble generator to the mineralization process with gypsum as the raw material [64] and used Venturi tubes with different diameters as mineralization reactors. Both reactors shortened the reaction time and produced nano-sized calcium carbonate. Altiner et al. [10] used a Venturi reactor with an ultrasonic probe to improve the agglomeration of calcium carbonate. Konopacka-Łyskawa et al. [94] used a bubble reactor to generate homogeneous vaterite calcium carbonate. Heping [95] used membrane electrolysis to achieve low-energy phosphogypsum mineralization.

4. Conclusions and Prospects

In summary, indirect mineralization, which uses by-product gypsum as a calcium source, can not only facilitate industrial “double waste” treatment but also produce high-value calcium carbonate with a higher raw material utilization rate and product purity than those of other processes. However, there are many difficulties and challenges too. On the one hand, reducing the capital cost and reagent consumption to make the entire CO2 mineralization process more economically viable. First of all, further development of leaching agents that with better cycle stability to simplify the process, such as aspartic acid, can not only facilitate the preparation of calcium carbonate products in a single-crystal form but also aid in the recovery and recycling of the leaching agents, which has both environmental and economic benefits. Secondly, coupling the indirect mineralization of by-product gypsum with the treatment of other industrial wastes will be a future development trend. On the other hand, the mineralization conditions for desulfurized gypsum and phosphogypsum are mild, whereas those for red gypsum are harsh. Therefore, understanding the internal differences in the mineralization of different gypsum by-products and providing clear information for selecting the best mineralization route will be an important future research direction. Meanwhile, an in-depth study of leaching and mineralization kinetics is required to understand indirect carbonation and provide basic data for its large-scale application.

Author Contributions

Conceptualization, B.W.; methodology, H.W.; software, B.W.; validation, C.L., Y.W. and Y.G.; formal analysis, Y.W.; investigation, B.W.; resources, B.W.; data curation, B.W.; writing—original draft preparation, B.W.; writing—review and editing, Y.W. and Y.G.; visualization, B.W.; supervision, C.L.; project administration, Y.W.; funding acquisition, C.L., Y.W. and Y.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Natural Science Foundation of Gansu Province, China, (Grant No. 21JR7RA226).

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, C.; Liang, C.; Chen, Z.; Di, Y.; Zheng, S.; Wei, S.; Sun, Z. Surface modification of calcium carbonate: A review of theories, methods and applications. J. Cent. South Univ. 2021, 28, 2589–2611. [Google Scholar] [CrossRef]
  2. Zhang, T.; Chu, G.; Lyu, J.; Cao, Y.; Xu, W.; Ma, K.; Song, L.; Yue, H.; Liang, B. CO2 mineralization of carbide slag for the production of light calcium carbonates. Chin. J. Chem. Eng. 2022, 43, 86–98. [Google Scholar] [CrossRef]
  3. Luo, J.; Kong, F.; Ma, X. Role of aspartic acid in the synthesis of spherical vaterite by the Ca(OH)2–CO2 reaction. Cryst. Growth Des. 2016, 16, 728–736. [Google Scholar] [CrossRef]
  4. Konopacka-Łyskawa, D. Synthesis methods and favorable conditions for spherical vaterite precipitation: A review. Crystals 2019, 9, 223. [Google Scholar] [CrossRef] [Green Version]
  5. Svenskaya, Y.I.; Fattah, H.; Inozemtseva, O.A.; Ivanova, A.G.; Shtykov, S.N.; Gorin, D.A.; Parakhonskiy, B.V. Key parameters for size- and shape-controlled synthesis of vaterite particles. Cryst. Growth Des. 2018, 18, 331–337. [Google Scholar] [CrossRef]
  6. Zeng, H.Y.; Yan, Z.L.; Jiao, M.R.; Xu, D.D.; Jiang, J.X. A novel method for preparing calcium carbonate particles: Thermal decomposition from calcium hydrogen carbonate solution. Key Eng. Mater. 2016, 697, 113–118. [Google Scholar]
  7. Yoo, Y.; Kim, I.; Lee, D.; Yong Choi, W.; Choi, J.; Jang, K.; Park, J.; Kang, D. Review of contemporary research on inorganic CO2 utilization via CO2 conversion into metal carbonate-based materials. J. Ind. Eng. Chem. 2022, 116, 60–74. [Google Scholar] [CrossRef]
  8. Lee, J.; Ryu, K.H.; Ha, H.Y.; Jung, K.-D.; Lee, J.H. Techno-economic and environmental evaluation of nano calcium carbonate production utilizing the steel slag. J. CO2 Util. 2020, 37, 113–121. [Google Scholar] [CrossRef]
  9. Luo, X.; Wei, C.; Li, X.; Deng, Z.; Li, M.; Fan, G. A green approach to prepare polymorph CaCO3 for clean utilization of salt gypsum residue and CO2 mineralization. Fuel 2023, 333, 126305. [Google Scholar] [CrossRef]
  10. Altiner, M.; Top, S.; Kaymakoğlu, B. Ultrasonic-assisted production of precipitated calcium carbonate particles from desulfurization gypsum. Ultrason Sonochem 2021, 72, 105421. [Google Scholar] [CrossRef]
  11. Wang, B.; Pan, Z.; Cheng, H.; Zhang, Z.; Cheng, F. A review of carbon dioxide sequestration by mineral Carbonation of Industrial Byproduct Gypsum. J. Clean. Prod. 2021, 302, 126930. [Google Scholar] [CrossRef]
  12. Owais, M.; Järvinen, M.; Taskinen, P.; Said, A. Experimental study on the extraction of calcium, magnesium, vanadium and silicon from steelmaking slags for improved mineral carbonation of CO2. J. CO2 Util. 2019, 31, 1–7. [Google Scholar] [CrossRef]
  13. Liu, S.; Liu, W.; Jiao, F.; Qin, W.; Yang, C. Production and resource utilization of flue gas desulfurized gypsum in China—A review. Environ. Pollut. 2021, 288, 117799. [Google Scholar] [CrossRef] [PubMed]
  14. Wu, F.; Ren, Y.; Qu, G.; Liu, S.; Chen, B.; Liu, X.; Zhao, C.; Li, J. Utilization path of bulk industrial solid waste: A review on the multi-directional resource utilization path of phosphogypsum. J. Environ. Manag. 2022, 313, 114957. [Google Scholar] [CrossRef] [PubMed]
  15. Ju, J.; Feng, Y.; Li, H.; Xu, C. Resource utilization of strongly acidic wastewater and red gypsum by a harmless self-treatment process. Process Saf. Environ. 2023, 172, 594–603. [Google Scholar] [CrossRef]
  16. Guan, Q.; Sui, Y.; Yu, W.; Bu, Y.; Zeng, C.; Liu, C.; Zhang, Z.; Gao, Z.; Chi, R. Deep removal of phosphorus and synchronous preparation of high-strength gypsum from phosphogypsum by crystal modification in NaCl-HCl solutions. Sep. Purif. Technol. 2022, 298, 121592. [Google Scholar] [CrossRef]
  17. Haneklaus, N.; Barbossa, S.; Basallote, M.D.; Bertau, M.; Bilal, E.; Chajduk, E.; Chernysh, Y.; Chubur, V.; Cruz, J.; Dziarczykowski, K.; et al. Closing the upcoming EU gypsum gap with phosphogypsum. Resour. Conserv. Recycl. 2022, 182, 106328. [Google Scholar] [CrossRef]
  18. Wu, Q.; Ma, H.; Chen, Q.; Huang, Z.; Zhang, C.; Yang, T. Preparation of waterproof block by silicate clinker modified FGD gypsum. Constr. Build. Mater. 2019, 214, 318–325. [Google Scholar] [CrossRef]
  19. Brooks, M.W.; Lynn, S. Recovery of calcium carbonate and hydrogen sulfide from waste calcium sulfide. Ind. Eng. Chem. Res. 1997, 36, 4236–4242. [Google Scholar] [CrossRef]
  20. de Beer, M.; Doucet, F.J.; Maree, J.P.; Liebenberg, L. Synthesis of high-purity precipitated calcium carbonate during the process of recovery of elemental sulphur from gypsum waste. Waste Manag. 2015, 46, 619–627. [Google Scholar] [CrossRef]
  21. Song, W.; Zhou, J.; Wang, B.; Li, S.; Han, J. New insight into investigation of reduction of desulfurization ash by pyrite for clean generation SO2. J. Clean. Prod. 2020, 253, 120026. [Google Scholar] [CrossRef]
  22. Zhang, X.; Wang, X.; Jin, B.; Liu, X.; Yang, L. Crystal structure formation of hemihydrate calcium sulfate whiskers (HH-CSWs) prepared using FGD gypsum. Polyhedron 2019, 173, 114140. [Google Scholar] [CrossRef]
  23. Liao, R.; Yu, H.; Lin, H.; Yang, P. A Quantitative study on three-dimensional pore parameters and physical properties of sodic soils restored by FGD gypsum and leaching water. J. Environ. Manag. 2019, 248, 109303. [Google Scholar] [CrossRef]
  24. Koralegedara, N.H.; Pinto, P.X.; Dionysiou, D.D.; Al-Abed, S.R. Recent advances in flue gas desulfurization gypsum processes and applications—A review. J. Environ. Manag. 2019, 251, 109572. [Google Scholar] [CrossRef]
  25. Zhu, X.; Zheng, S.; Zhang, Y.; Fang, Z.Z.; Zhang, M.; Sun, P.; Li, Q.; Zhang, Y.; Li, P.; Jin, W. Potentially more ecofriendly chemical pathway for production of high-purity TiO2 from titanium slag. Acs Sustain. Chem. Eng. 2019, 7, 4821–4830. [Google Scholar] [CrossRef]
  26. Pyagai, I.; Zubkova, O.; Babykin, R.; Toropchina, M.; Fediuk, R. Influence of impurities on the process of obtaining calcium carbonate during the processing of phosphogypsum. Materials 2022, 15, 4335. [Google Scholar] [CrossRef]
  27. Wang, Y.; Song, L.; Ma, K.; Liu, C.; Tang, S.; Yan, Z.; Yue, H.; Liang, B. an integrated absorption–mineralization process for CO2 capture and sequestration: Reaction mechanism, recycling stability, and energy evaluation. ACS Sustain. Chem. Eng. 2021, 9, 16577–16587. [Google Scholar] [CrossRef]
  28. Baena-Moreno, F.M.; le Saché, E.; Hurd Price, C.A.; Reina, T.R.; Navarrete, B. From biogas upgrading to CO2 utilization and waste recycling: A novel circular economy approach. J. CO2 Util. 2021, 47, 101496. [Google Scholar] [CrossRef]
  29. Rahmani, O. An experimental study of accelerated mineral carbonation of industrial waste red gypsum for CO2 sequestration. J. CO2 Util. 2020, 35, 265–271. [Google Scholar] [CrossRef]
  30. Xie, H.; Yue, H.; Zhu, J.; Liang, B.; Li, C.; Wang, Y.; Xie, L.; Zhou, X. Scientific and engineering progress in CO2 mineralization using industrial waste and natural minerals. Engineering 2015, 1, 150–157. [Google Scholar] [CrossRef] [Green Version]
  31. Meng, J.; Liao, W.; Zhang, G. Emerging CO2-mineralization technologies for co-utilization of industrial solid waste and carbon resources in China. Minerals 2021, 11, 274. [Google Scholar] [CrossRef]
  32. Li, Y.; Ni, W.; Duan, P.; Zhang, S.; Wang, J. Experimental study and mechanism analysis of preparation of alpha-calcium sulfate hemihydrate from FGD gypsum with dynamic method. Materials 2022, 15, 3382. [Google Scholar] [CrossRef] [PubMed]
  33. Shao, S.; Ma, B.; Wang, X.; Zhang, W.; Chen, Y.; Wang, C. Nitric acid pressure leaching of limonitic laterite ores: Regeneration of HNO3 and Simultaneous Synthesis of Fibrous CaSO4·2H2O by-products. J. Cent. South Univ. 2020, 27, 3249–3258. [Google Scholar] [CrossRef]
  34. Zhang, S.; Zhao, Y.; Ding, H.; Qiu, J.; Guo, Z. Recycling flue gas desulfurisation gypsum and phosphogypsum for cemented paste backfill and its acid resistance. Constr. Build Mater. 2021, 275, 122170. [Google Scholar] [CrossRef]
  35. Wang, B.; Pan, Z.; Du, Z.; Cheng, H.; Cheng, F. Effect of impure components in flue gas desulfurization (FGD) gypsum on the generation of polymorph CaCO3 during carbonation reaction. J. Hazard. Mater. 2019, 369, 236–243. [Google Scholar] [CrossRef]
  36. Tokpayev, R.; Khavaza, T.; Ibraimov, Z.; Kishibayev, K.; Atchabarova, A.; Abdimomyn, S.; Abduakhytova, D.; Nauryzbayev, M. Phosphogypsum conversion under conditions of SC-CO2. J. CO2 Util. 2022, 63, 102120. [Google Scholar] [CrossRef]
  37. Peng, X.; Zheng, J.; Liu, Q.; Hu, Q.; Sun, X.; Li, J.; Liu, W.; Lin, Z. Efficient removal of iron from red gypsum via synergistic regulation of gypsum phase transformation and iron speciation. Sci. Total Environ. 2021, 791, 148319. [Google Scholar] [CrossRef]
  38. Wu, F.; Zhao, C.; Qu, G.; Liu, S.; Ren, Y.; Chen, B.; Li, J.; Liu, L. A critical review of the typical by-product clean ecology links in the Chinese phosphorus chemical industry in China: Production technologies, Fates and Future Directions. J. Environ. Chem. Eng. 2022, 10, 106685. [Google Scholar] [CrossRef]
  39. Gázquez, M.J.; Bolívar, J.P.; Garcia-Tenorio, R.; Vaca, F. A review of the production cycle of titanium dioxide pigment. Mater. Sci. Appl. 2014, 5, 441–458. [Google Scholar] [CrossRef] [Green Version]
  40. Leiva, C.; García Arenas, C.; Vilches, L.F.; Vale, J.; Gimenez, A.; Ballesteros, J.C.; Fernández-Pereira, C. Use of FGD gypsum in fire resistant panels. Waste Manag. 2010, 30, 1123–1129. [Google Scholar] [CrossRef]
  41. Altiner, M. Effect of alkali types on the production of calcium carbonate particles from gypsum waste for fixation of CO2 by mineral carbonation. Int. J. Coal Prep. Util. 2019, 39, 113–131. [Google Scholar] [CrossRef]
  42. Ma, Y.; Nie, Q.; Xiao, R.; Hu, W.; Han, B.; Polaczyk, P.A.; Huang, B. Experimental investigation of utilizing waste flue gas desulfurized gypsum as backfill materials. Constr. Build Mater. 2020, 245, 118393. [Google Scholar] [CrossRef]
  43. Nizevičienė, D.; Vaičiukynienė, D.; Michalik, B.; Bonczyk, M.; Vaitkevičius, V.; Jusas, V. The treatment of phosphogypsum with zeolite to use it in binding material. Constr. Build Mater. 2018, 180, 134–142. [Google Scholar] [CrossRef]
  44. Romero-Hermida, M.I.; Borrero-López, A.M.; Flores-Alés, V.; Alejandre, F.J.; Franco, J.M.; Santos, A.; Esquivias, L. Characterization and analysis of the carbonation process of a lime mortar obtained from phosphogypsum waste. Int. J. Environ. Res. Public Health 2021, 18, 6664. [Google Scholar] [CrossRef] [PubMed]
  45. Rahmani, O. Siderite precipitation using by-product red gypsum for CO2 sequestration. J. CO2 Util. 2018, 24, 321–327. [Google Scholar] [CrossRef]
  46. Marian, N.M.; Perotti, M.; Indelicato, C.; Magrini, C.; Giorgetti, G.; Capitani, G.; Viti, C. From high-volume industrial waste to new ceramic material: The case of red gypsum muds in the TiO2 industry. Ceram. Int. 2023, 49, 15034–15043. [Google Scholar] [CrossRef]
  47. Zeng, C.; Hu, H.; Feng, X.; Chen, M.; Shi, Q.; Chen, M.; Zhang, Q. Efficient removal of lead impurity for the purification and recycling of nickel from secondary sources based on ball-milling activated CaCO3. J. Environ. Chem. Eng. 2021, 9, 106737. [Google Scholar] [CrossRef]
  48. Lian, F.; Gao, S.; Fu, Q.; Wu, Y.; Wang, J.; Huang, Q.; Hu, S. A comprehensive study of phosphorus removal and recovery with a Fe-loaded sulfoaluminate cement (FSC) adsorbent. J. Water Process Eng. 2021, 39, 101744. [Google Scholar] [CrossRef]
  49. O’Connor, W.K.; Dahlin, D.C.; Nilsen, D.N.; Gerdemann, S.J.; Rush, G.E.; Penner, L.R.; Walters, R.P.; Turner, P.C. Continuing Studies on Direct Aqueous Mineral Carbonation of CO2 Sequestration; USDOE Office of Fossil Energy (FE) (US): Washington, DC, USA, 2002.
  50. Romanov, V.; Soong, Y.; Carney, C.; Rush, G.E.; Nielsen, B.; O’Connor, W. Mineralization of carbon dioxide: A literature review. ChemBioEng Rev. 2015, 2, 231–256. [Google Scholar] [CrossRef]
  51. Cárdenas-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] [Green Version]
  52. Part, W.K.; Ramli, M.; Cheah, C.B. An overview on the influence of various factors on the properties of geopolymer concrete derived from industrial by-products. Constr. Build Mater. 2015, 77, 370–395. [Google Scholar] [CrossRef]
  53. Baojun, Y.; Mengmeng, Y.; Bainian, W.; Xiaoyu, F.; Qiang, W. A new route to synthesize calcium carbonate microspheres from phosphogypsum. Mater. Res. Express 2019, 6, 045042. [Google Scholar] [CrossRef]
  54. Chen, Q.; Ding, W.; Sun, H.; Peng, T.; Ma, G. Indirect mineral carbonation of phosphogypsum for CO2 sequestration. Energy 2020, 206, 118148. [Google Scholar] [CrossRef]
  55. Gong, Y.; Zhu, X.; Yang, Z.; Zhang, X.; Li, C. Indirect aqueous carbonation of CaSO4·2H2O with aspartic acid as a recyclable additive. RSC Adv. 2022, 12, 26556–26564. [Google Scholar] [CrossRef] [PubMed]
  56. Blencoe, J.G.; Palmer, D.A.; Anovitz, L.M.; Beard, J.S. Carbonation of metal silicates for long-term CO2 sequestration 2020. Available online: https://www.freepatentsonline.com/10632418.html (accessed on 12 April 2023).
  57. Pérez-Moreno, S.M.; Gázquez, M.J.; Bolívar, J.P. CO2 sequestration by indirect carbonation of artificial gypsum generated in the manufacture of titanium dioxide pigments. Chem. Eng. J. 2015, 262, 737–746. [Google Scholar] [CrossRef]
  58. Romero-Hermida, I.; Santos, A.; Pérez-López, R.; García-Tenorio, R.; Esquivias, L.; Morales-Flórez, V. New method for carbon dioxide mineralization based on phosphogypsum and aluminium-rich industrial wastes resulting in valuable carbonated by-products. J. CO2 Util. 2017, 18, 15–22. [Google Scholar] [CrossRef] [Green Version]
  59. Rahmani, O. CO2 sequestration by indirect mineral carbonation of industrial waste red gypsum. J. CO2 Util. 2018, 27, 374–380. [Google Scholar] [CrossRef]
  60. Altiner, M.; Top, S.; Kaymakoğlu, B.; Seçkin, İ.Y.; Vapur, H. Production of precipitated calcium carbonate particles from gypsum waste using venturi tubes as a carbonation zone. J. CO2 Util. 2019, 29, 117–125. [Google Scholar] [CrossRef]
  61. Xue, S.; Li, M.; Jiang, J.; Millar, G.J.; Li, C.; Kong, X. Phosphogypsum stabilization of bauxite residue: Conversion of its alkali characteristics. J. Environ. Sci. 2019, 77, 1–10. [Google Scholar] [CrossRef]
  62. Biyoune, M.G.; Bouargane, B.; Idboufrade, A.; Marrouche, A.; Atbir, A.; Boukbir, L.; Mançour-billah, S. New procedure for water-salinity reduction using phosphogypsum waste and carbon dioxide resulting in useful compounds formation. Nanotechnol. Environ. Eng. 2021, 6, 33. [Google Scholar] [CrossRef]
  63. Kelly, K.E.; Silcox, G.D.; Sarofim, A.F.; Pershing, D.W. An evaluation of ex situ, industrial-scale, aqueous CO2 mineralization. Int. J. Greenh Gas. Con. 2011, 5, 1587–1595. [Google Scholar] [CrossRef]
  64. Kim, G.; Kim, S.; Kim, M.-J. Effect of sucrose on CO2 storage, vaterite content, and caco3 particle size in indirect carbonation using seawater. J. CO2 Util. 2022, 57, 101894. [Google Scholar] [CrossRef]
  65. Azdarpour, A.; Asadullah, M.; Mohammadian, E.; Hamidi, H.; Junin, R.; Karaei, M.A. A review on carbon dioxide mineral carbonation through pH-swing process. Chem. Eng. J. 2015, 279, 615–630. [Google Scholar] [CrossRef]
  66. Rim, G.; Roy, N.; Zhao, D.; Kawashima, S.; Stallworth, P.; Greenbaum, S.G.; Park, A.-H.A. CO2 utilization in built environment via the PCO2 swing carbonation of alkali solid wastes with different mineralogy. Faraday Discuss. 2021, 230, 187–212. [Google Scholar] [CrossRef]
  67. Park, A.-H.A.; Jadhav, R.; Fan, L.-S. CO2 mineral sequestration: Chemically enhanced aqueous carbonation of serpentine. Can. J. Chem. Eng. 2003, 81, 885–890. [Google Scholar] [CrossRef]
  68. Rahmani, O.; Tyrer, M.; Junin, R. Calcite precipitation from by-product red gypsum in aqueous carbonation process. RSC Adv. 2014, 4, 45548–45557. [Google Scholar] [CrossRef]
  69. Azdarpour, A.; Asadullah, M.; Junin, R.; Mohammadian, E.; Hamidi, H.; Daud, A.R.M.; Manan, M. Extraction of calcium from red gypsum for calcium carbonate production. Fuel Process. Technol. 2015, 130, 12–19. [Google Scholar] [CrossRef]
  70. Azdarpour, A.; Asadullah, M.; Mohammadian, E.; Junin, R.; Hamidi, H.; Manan, M.; Daud, A.R.M. Mineral carbonation of red gypsum via pH-swing process: Effect of CO2 pressure on the efficiency and products characteristics. Chem. Eng. J. 2015, 264, 425–436. [Google Scholar] [CrossRef]
  71. Azdarpour, A.; Karaei, M.A.; Hamidi, H.; Mohammadian, E.; Barati, M.; Honarvar, B. CO2 sequestration using red gypsum via pH-swing process: Effect of carbonation temperature and NH4HCO3 on the process efficiency. Int. J. Miner. Process. 2017, 169, 27–34. [Google Scholar] [CrossRef] [Green Version]
  72. Ding, W.; Chen, Q.; Sun, H.; Peng, T. Modified phosphogypsum sequestrating CO2 and characteristics of the carbonation product. Energy 2019, 182, 224–235. [Google Scholar] [CrossRef]
  73. Azdarpour, A.; Afkhami Karaei, M.; Hamidi, H.; Mohammadian, E.; Honarvar, B. CO2 sequestration through direct aqueous mineral carbonation of red gypsum. Petroleum 2018, 4, 398–407. [Google Scholar] [CrossRef]
  74. Azdarpour, A.; Asadullah, M.; Junin, R.; Manan, M.; Hamidi, H.; Mohammadian, E. Direct carbonation of red gypsum to produce solid carbonates. Fuel Process. Technol. 2014, 126, 429–434. [Google Scholar] [CrossRef]
  75. Wenjin, D.; Chen, Q.; Hongjuan, S.; Tongjiang, P.; Peng, T. Modified mineral carbonation of phosphogypsum for CO2 sequestration. J. CO2 Util. 2019, 34, 507–515. [Google Scholar]
  76. Chen, Q.; Ding, W.; Sun, H.; Peng, T.; Ma, G. Utilization of phosphogypsum to prepare high-purity CaCO3 in the NH4Cl–NH4OH–CO2 system. Acs Sustain. Chem. Eng. 2020, 8, 11649–11657. [Google Scholar] [CrossRef]
  77. Ding, W.; Qiao, J.; Zeng, L.; Sun, H.; Peng, T. Desulfurization gypsum carbonation for CO2 sequestration by using recyclable ammonium salt. Int. J. Greenh Gas Con. 2023, 123, 103843. [Google Scholar] [CrossRef]
  78. Fredd, C.N.; Fogler, H.S. The influence of chelating agents on the kinetics of calcite dissolution. J. Colloid Interface Sci. 1998, 204, 187–197. [Google Scholar] [CrossRef]
  79. Nowack, B. Environmental chemistry of aminopolycarboxylate chelating agents. Environ. Sci. Technol. 2002, 36, 4009–4016. [Google Scholar] [CrossRef]
  80. Ghoorah, M.; Dlugogorski, B.Z.; Balucan, R.D.; Kennedy, E.M. Selection of acid for weak acid processing of wollastonite for mineralisation of CO2. Fuel 2014, 122, 277–286. [Google Scholar] [CrossRef]
  81. Kim, M.-J.; Jeon, J. Effects of Ca-ligand stability constant and chelating agent concentration on the CO2 storage using paper sludge ash and chelating agent. J. CO2 Util. 2020, 40, 101202. [Google Scholar] [CrossRef]
  82. Yuan, X.; Chen, X.; Gao, S.; Wang, Y.; Yang, L.; Zhang, Q.; Chen, Y.; Wang, B.; Yang, B. Preparation of calcium carbonate microrods from the gypsum scale layer of evaporation equipment. RSC Adv. 2022, 12, 10584–10591. [Google Scholar] [CrossRef]
  83. Huang, W.; Wang, Q.; Chi, W.; Cai, M.; Wang, R.; Fu, Z.; Xie, J.; Zou, Z. Multiple crystallization pathways of amorphous calcium carbonate in the presence of poly(aspartic acid) with a chain length of 30. CrystEngComm 2022, 24, 4809–4818. [Google Scholar] [CrossRef]
  84. Liu, X.; Wang, B.; Zhang, Z.; Pan, Z.; Cheng, H.; Cheng, F. Glycine-induced synthesis of vaterite by direct aqueous mineral carbonation of desulfurization gypsum. Environ. Chem. Lett. 2022, 20, 2261–2269. [Google Scholar] [CrossRef]
  85. 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]
  86. Zheng, X.; Zhang, L.; Feng, L.; He, Q.; Ji, L.; Yan, S. Insights into dual functions of amino acid salts as CO2 Carriers and CaCO3 regulators for integrated CO2 absorption and mineralisation. J. CO2 Util. 2021, 48, 101531. [Google Scholar] [CrossRef]
  87. 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]
  88. Contreras, M.; Pérez-López, R.; Gázquez, M.J.; Morales-Flórez, V.; Santos, A.; Esquivias, L.; Bolívar, J.P. Fractionation and fluxes of metals and radionuclides during the recycling process of phosphogypsum wastes applied to mineral CO₂ sequestration. Waste Manag. 2015, 45, 412–419. [Google Scholar] [CrossRef] [Green Version]
  89. Melnikov, S.S.; Mugtamov, O.A.; Zabolotsky, V.I. Study of electrodialysis concentration process of inorganic acids and salts for the two-stage conversion of salts into acids utilizing bipolar electrodialysis. Sep. Purif. Technol. 2020, 235, 116198. [Google Scholar] [CrossRef]
  90. van Linden, N.; Bandinu, G.L.; Vermaas, D.A.; Spanjers, H.; van Lier, J.B. Bipolar membrane electrodialysis for energetically competitive ammonium removal and dissolved ammonia production. J. Clean. Prod. 2020, 259, 120788. [Google Scholar] [CrossRef]
  91. 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]
  92. Kuldeep; Badenhorst, W.D.; Kauranen, P.; Pajari, H.; Ruismäki, R.; Mannela, P.; Murtomäki, L. Bipolar membrane electrodialysis for sulfate recycling in the metallurgical industries. Membranes 2021, 11, 718. [Google Scholar] [CrossRef]
  93. Luo, Y.; Liu, Y.; Shen, J.; Van der Bruggen, B. Application of bipolar membrane electrodialysis in environmental protection and resource recovery: A review. Membranes 2022, 12, 829. [Google Scholar] [CrossRef] [PubMed]
  94. Donata Konopacka-Łyskawa; Barbara Kościelska; Marcin łapiński precipitation of spherical vaterite particles via carbonation route in the bubble column and the gas-lift reactor. Jom-Us 2019, 71, 1041–1048. [CrossRef] [Green Version]
  95. Xie, H.; Wang, J.; Hou, Z.; Wang, Y.; Liu, T.; Tang, L.; Jiang, W. CO2 sequestration through mineral carbonation of waste phosphogypsum using the technique of membrane electrolysis. Environ. Earth Sci 2016, 75, 1216. [Google Scholar] [CrossRef]
Figure 1. SEM images of industrial by-product gypsum: (a) desulfurization gypsum (Reprinted with permission from Ref. [35] 2019, Wang et al.), (b) phosphogypsum (Reprinted with permission from Ref. [36] 2022, Tokpayev et al.), and (c) red gypsum (Reprinted with permission from Ref. [37] 2021, Peng et al.).
Figure 1. SEM images of industrial by-product gypsum: (a) desulfurization gypsum (Reprinted with permission from Ref. [35] 2019, Wang et al.), (b) phosphogypsum (Reprinted with permission from Ref. [36] 2022, Tokpayev et al.), and (c) red gypsum (Reprinted with permission from Ref. [37] 2021, Peng et al.).
Sustainability 15 09629 g001
Figure 2. Indirect mineralization of by-product gypsum.
Figure 2. Indirect mineralization of by-product gypsum.
Sustainability 15 09629 g002
Figure 3. Indirect mineralization process of phosphogypsum by alkali leaching (Reprinted with permission from Ref. [51] 2011, Cárdenas-Escudero et al.).
Figure 3. Indirect mineralization process of phosphogypsum by alkali leaching (Reprinted with permission from Ref. [51] 2011, Cárdenas-Escudero et al.).
Sustainability 15 09629 g003
Figure 4. Schematic diagram of the indirect mineralization of red gypsum using the pH swing process (Reprinted with permission from Ref. [68] 2014, Rahmani et al.).
Figure 4. Schematic diagram of the indirect mineralization of red gypsum using the pH swing process (Reprinted with permission from Ref. [68] 2014, Rahmani et al.).
Sustainability 15 09629 g004
Figure 5. Effect of carbonation reaction temperature on XRD pattern (a) and morphology (b) of products. Influence of temperature on the SEM images of carbonation product: (c) 20 °C, (d) 30 °C, (e) 40 °C, (f) 50 °C, (g) 60 °C, (h) 80 °C and (i) 100 °C (Reprinted with permission from Ref. [72] 2019, Ding et al.).
Figure 5. Effect of carbonation reaction temperature on XRD pattern (a) and morphology (b) of products. Influence of temperature on the SEM images of carbonation product: (c) 20 °C, (d) 30 °C, (e) 40 °C, (f) 50 °C, (g) 60 °C, (h) 80 °C and (i) 100 °C (Reprinted with permission from Ref. [72] 2019, Ding et al.).
Sustainability 15 09629 g005
Figure 6. (a) Change in leaching efficiency with cycle time when NH4Cl is the leaching agent; (b) XRD patterns of carbonated products with various cycle numbers. (Reprinted with permission from Ref. [76] 2014, Rahmani et al.).
Figure 6. (a) Change in leaching efficiency with cycle time when NH4Cl is the leaching agent; (b) XRD patterns of carbonated products with various cycle numbers. (Reprinted with permission from Ref. [76] 2014, Rahmani et al.).
Sustainability 15 09629 g006
Figure 7. Schematic diagram of carbonation and carbon sequestration of phosphogypsum (Reprinted with permission from Ref. [75] 2019, Ding et al.).
Figure 7. Schematic diagram of carbonation and carbon sequestration of phosphogypsum (Reprinted with permission from Ref. [75] 2019, Ding et al.).
Sustainability 15 09629 g007
Figure 8. (a) Flow chart of the indirect mineralization of CaSO4 2H2O with aspartic acid as a leaching agent; (b) variations in the pH and conductivity of the reaction system over time (Areas A-C are the induction, growth and stabilization period of CaCO3 formation, respectively); (c) XRD and SEM images of the obtained CaCO3 products (Reprinted with permission from Ref. [55] 2022, Gong et al.).
Figure 8. (a) Flow chart of the indirect mineralization of CaSO4 2H2O with aspartic acid as a leaching agent; (b) variations in the pH and conductivity of the reaction system over time (Areas A-C are the induction, growth and stabilization period of CaCO3 formation, respectively); (c) XRD and SEM images of the obtained CaCO3 products (Reprinted with permission from Ref. [55] 2022, Gong et al.).
Sustainability 15 09629 g008
Figure 9. (a) Changes in the leaching capacity and recovery efficiency of aspartic acid leaching cycle; (b) XRD patterns and SEM images of CaCO3 recorded during the cycling (Reprinted with permission from Ref. [55] 2022, Gong et al.).
Figure 9. (a) Changes in the leaching capacity and recovery efficiency of aspartic acid leaching cycle; (b) XRD patterns and SEM images of CaCO3 recorded during the cycling (Reprinted with permission from Ref. [55] 2022, Gong et al.).
Sustainability 15 09629 g009
Figure 10. Schematic diagram of bipolar membrane electrodialysis application in phosphogypsum mineralization.
Figure 10. Schematic diagram of bipolar membrane electrodialysis application in phosphogypsum mineralization.
Sustainability 15 09629 g010
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

Wu, B.; Wang, H.; Li, C.; Gong, Y.; Wang, Y. Progress in the Preparation of Calcium Carbonate by Indirect Mineralization of Industrial By-Product Gypsum. Sustainability 2023, 15, 9629. https://0-doi-org.brum.beds.ac.uk/10.3390/su15129629

AMA Style

Wu B, Wang H, Li C, Gong Y, Wang Y. Progress in the Preparation of Calcium Carbonate by Indirect Mineralization of Industrial By-Product Gypsum. Sustainability. 2023; 15(12):9629. https://0-doi-org.brum.beds.ac.uk/10.3390/su15129629

Chicago/Turabian Style

Wu, Baizhi, Haibin Wang, Chunlei Li, Yuan Gong, and Yi Wang. 2023. "Progress in the Preparation of Calcium Carbonate by Indirect Mineralization of Industrial By-Product Gypsum" Sustainability 15, no. 12: 9629. https://0-doi-org.brum.beds.ac.uk/10.3390/su15129629

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