Next Article in Journal
Battery Thermal Management: An Application to Petrol Hybrid Electric Vehicles
Previous Article in Journal
Green Practices in Mega Development Projects of China–Pakistan Economic Corridor
Previous Article in Special Issue
Implementation of Magnetic Nanostructured Adsorbents for Heavy Metals Separation from Textile Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Remediation of Micro-Pollution in an Alkaline Washing Solution of Fly Ash Using Simulated Exhaust Gas: Parameters and Mechanism

1
School of Environmental Science and Engineering, Nanjing Tech University, Nanjing 211816, China
2
State Environmental Protection Key Laboratory of Pesticide Environmental Assessment and Pollution Control, Nanjing Institute of Environmental Sciences, Ministry of Ecology and Environment (MEE), Nanjing 210042, China
3
The Key Laboratory of Water and Air Pollution Control of Guangdong Province, South China Institute of Environmental Sciences, Ministry of Ecology and Environment (MEE), Guangzhou 510655, China
4
Department of Inorganic Chemistry, Faculty of Chemistry, Maria Curie-Skłodowska University, Maria Curie-Skłodowska Sq. 2, 20-031 Lublin, Poland
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Sustainability 2023, 15(7), 5873; https://doi.org/10.3390/su15075873
Submission received: 2 February 2023 / Revised: 5 March 2023 / Accepted: 21 March 2023 / Published: 28 March 2023
(This article belongs to the Special Issue Advances in Sustainable Treatment of Complex Wastewater)

Abstract

:
Currently, there is an urgent need to remediate heavy metals (HMs) and high alkalinity in the washing solution of fly ash (FA). This study investigated the remediation with simulated exhaust gases of two CO2 partial pressure and revealed the removal efficiency of target pollutants, mainly including Pb ions. The results verify that under the preferred conditions of 25 °C and 15 mL/min flow rate, bubbling two kinds of simulated flue gases could efficiently remove 97.9–99.2% of Pb ions. Moreover, the initial 40 min removal of Pb ions fits in a way with a pseudo-first-order equation. Based on the thermodynamic parameters, we infer that the removal of Pb ions was a spontaneous, exothermic, and entropy-decreasing process. Furthermore, residual HMs and terminal pH after remediation of the FA washing solution basically met the regulatory threshold values of the integrated wastewater discharge standard in China (GB 8978−1996). Additionally, the particles obtained from the washing solution of FA were identified as CaCO3, which was mainly composed of vaterite and calcite crystalline. This study provides a fundamental guide for remediating multiple pollutants in the washing solution of FA and simultaneously sequestrating carbon emissions from power plants and industries.

1. Introduction

During rapid urbanization, municipal solid waste incineration (MSWI) has gradually and globally been selected as the most successful strategy for waste management [1,2,3]. However, MSWI could generate considerable hazardous alkaline fly ash (FA), which has been listed in the Chinese National Hazardous Waste Inventory (No. HW18). Presently, the stabilization/solidification of heavy metals (HMs) and the detoxification of dioxin are the main research hotspots for FA treatment. Although alkaline materials can sequestrate CO2 gas through mineral carbonation of alkaline components [4], excessive alkaline FA is seldom applied to reduce the greenhouse gas CO2 and acid exhaust gas emission [5,6]. In particular, global CO2 emissions from coal plants alone have been reported to reach 10.5 Gt in 2021 [7]. Therefore, it is significant to simultaneously reduce the pollution of FA and carbon emissions from flue gases.
As one of the typical alkaline industrial wastes, FA contains a large amount of calcium oxide (CaO), indicating that FA could capture CO2 with the purpose of ‘using waste to treat waste’. According to the available documents, direct carbonation of FA with CO2 during the gas/solid reactions could achieve transient mass transfer of CO2 and partial immobilization of HMs [8,9,10]. It is difficult, however, to control the carbonation ratio under conditions of high pressure and high temperature [11,12]. Moreover, Dananjayan et al. reported that in the case of the wet carbonation of coal fly ash (CFA), adoption with a water slurry of 15 as L/S ratio could achieve the maximum sequestration capacity as 50.3 g CO2/kg waste, compared to the maximum sequestration capacity of 26.3 g CO2/kg dry CFA for dry carbonation [13]. This indicates that adding H2O into the FA carbonation process could significantly increase the transfer of CO2 during the slow route of dry carbonation. Thus, the carbonation of FA is mainly carried out under aqueous conditions, dominated by alkaline solution.
Wet carbonation is a major process to simultaneously achieve the washing pretreatment of FA and in controlling CO2 emission, which is favored by researchers because it exhibits a wide range of advantages: simple equipment, lower energy consumption, higher carbonation efficiency, and the desalination of FA [14,15]. As for the wet carbonation reactions, much attention has been focused on the effects of particle size [16], temperature [17], pressure, liquid–solid ratio [4], time, CO2 volume fraction and flow rate, pH and acid gas (e.g., SO2 [18]). To recycle carbonates with high purity [19], various leaching agents, including water, acids [20,21,22], alkali [23] and salts [24,25,26,27], have been applied to accelerate the extraction of calcium components from alkaline waste, followed by the carbonation under two-phase gas–liquid conditions.
Our prior experiments demonstrated that washing FA could promote the dissolution of hazardous materials into the washing solution. Then, bubbling CO2 into the aforementioned FA washing solution can effectively deposit the amphoteric HMs, including Pb, Zn, and Cu [28], and control the high alkalinity [29]. In addition, direct CO2 capture using alkaline FA has another limitation, i.e., this process does not destroy residual dioxins and may release CO2 during further pyrolysis of the carbonized FA. Our prior study used a microwave (MW)-induced pyrolysis reaction to treat FA, which had been pre-treated with phosphate-containing anionic water solution [29,30]. The results demonstrated that above 99% of dioxin in FA was effectively degraded within the pyrolysis temperature varying between 380−610 °C [31]. Therefore, we propose that the detoxification of FA should be treated in steps. For example, soluble contaminants and dioxin in FA can be separately treated by a washing pre-treatment and an MW-induced pyrolysis process.
As for the washing solution of FA, bubbling CO2 exhaust gas is expected to remediate the micro-pollution and sequestrate CO2 simultaneously. Previously, high removal efficiencies (up to 99.37–99.69%) of Zn(II), Pb(II), Cu(II), and Cr(III) have been observed in wastewater treated by fly ash–lime carbonation [32]. Furthermore, a high concentration of Ca(OH)2(aq) was prone to generating CaCO3 [33] when CO2 was bubbled into alkali wastewater. Meanwhile, the calcite precipitation reaction could effectively reduce the pH of alkali wastewater (e.g., CO2 + 2OH = CO32− + H2O; Ca2+ + CO32− = CaCO3; ∆H298K = −178 kJ/mol [34]). The simultaneous removal mechanism of amphoteric HMs during the recycling CaCO3, however, has not been clearly elucidated due to low solubility (e.g., Ksp(PbCO3) = 3.3 × 10−14, Ksp(CaCO3) = 2.7 × 10−9) and complex components. Additionally, based on the available data, exhaust emissions from power plant/industrial emissions which contain CO2 have been rarely applied to remediate the micro-pollution in the washing solution of FA.
In this study, the main objectives were to remediate target HMs (e.g., Pb, Cu, and Zn), neutralize the alkalinity of the washing solution of FA, and sequestrate CO2 using simulated power plant/industrial emissions. We simulated dry-based exhaust gases of the coal-fired power plant (CFPP) and the cement/steel industry with 15% and 33% partial pressure of CO2 [35], respectively. Then, we remediated the washing solution of FA with the simulated exhaust gases and revealed variations of target HM concentrations, including Pb ions and pH values versus reaction time. Furthermore, we investigated the thermodynamics and mechanism of Pb removal. Eventually, the obtained particles were characterized in detail. This study provides fundamental guidance for the simultaneous remediation of micro-pollution in the washing solution of FA and the reduction of carbon emissions in power plants/industries.

2. Materials and Methods

2.1. Fly Ash Sample and Analysis Testing

Fly ash used in this study was collected from a MSWI facility in Suzhou, China. The furnace was equipped with an air pollution control system composed mainly of a semi-dry flue gas cleaning tower, an active carbon absorption reactor, and a bag filter. Fly ash was homogeneously mixed, further dried at 105 °C for 24 h, and eventually characterized for its chemical composition. The major elements of the original fly ash were characterized with X-ray fluorescence (XRF-1700, Shimadzu Corporation, Kyoto, Japan). XRD patterns were recorded using Cu Kα radiation (e.g., 50 KeV, 200 mA; an X’Pert Pro diffractometer) to identify the crystal phases of the original fly ash and the precipitates (Smart Lab, Rigaku, Tokyo, Japan). Next, the morphology and composition of fly ash were characterized with field emission scanning electron microscope-energy dispersion X-ray spectroscopy (SEM-EDS, Regulus8100, Hitachi, Tokyo, Japan). The samples were loaded with gold before examination by a JSM6301 analyzer with a voltage of 15 kV and 1.5 nm resolution. The detailed characterization of the original fly ash was discussed in our previous study [33].

2.2. Bubbling Experiments for the Washing Solution of Fly Ash

In this study, the washing parameters of FA have been mainly selected as liquid-to-solid (L/S) ratio of 3–10 mL/g, 1 L of washing solution, 10 min of washing duration, and standard ambient temperature and pressure. After the washing process, the leachate was separated with vacuum filtration. The residual HMs and pH values in the leachate were determined using the inductively coupled plasma mass spectrometer (ICP-MS, iCAP RQ, Thermo Fisher, Waltham, MA, USA) or inductively coupled plasma optical emission spectrometer (ICP-OES, iCAP 7200, Thermo Fisher, Waltham, MA, USA) and pH meter (PHC101, HACH, Loveland, CO, USA). Finally, the washed FA was dried with a freeze-drying process (−80 °C for 24 h) for further characterization. Moreover, the industrial tail gases from the CFPP and cement/steel industries with CO2 and N2 (CO2 partial pressure as 15% and 33%, respectively) were simulated, and the effects of gas partial pressures on element removals in the washing solution were compared. The whole apparatus is available in Figure S1. The details of the bubbling experiments were described as follows:
First, high-purity N2 and high-purity CO2 were connected with an electromagnetic mass flowmeter (LZB-3WB, Senlod, Nanjing, China) to control the flow rates of the two gases proportionally. Second, a gas mixing device was applied to mix thoroughly the aforementioned gases, which were further heated to the specified temperatures (e.g., 25 °C, 50 °C, and 80 °C) through a gas preheating device. Additionally, the flow rates of 10 mL/min (sccm) and 15 sccm were conducted to reveal the influence of gas flow rates on the treatment of the washing solution.
The washing solutions of FA were subjected to the bubbling system using sufficiently mixed and heated gases. The FA washing solution was placed into a special reaction vessel, in which a pH meter (continuous measurement mode) was immersed and a temperature sensor was connected to ensure a constant temperature of 25 °C. Then, the fully mixed gas of 25 °C was passed into the aforementioned reactor. When the pH value changed to the preset value (pH = 8), the supply of the simulated gas was stopped. The quantitative suspensions were removed from the reactor, accompanied by a recording of the pH and the bubbling time. The suspensions were filtered through 0.45 μm filters, and the filtrates were further collected into 50 mL colorimetric tubes and acidified (final pH value < 2). The concentrations of residual HMs were determined by ICP-OES, and the minor elements were quantified by ICP-MS. The above experiments were also conducted at 50 °C and 80 °C (heated gas and thermostatic water bath), respectively.
To compare the characteristics of the particles precipitated at different flow rates of gases, the bubbling processes for gas flow rates of 10, 15, 30, 45, and 61 sccm were studied. Since the partial pressure of CO2 was set to 33%, the bubbling experiments using pure CO2 with flow rates of 3, 4.95, 9.9, 14.85, and 20 sccm were re-perfomed. The process of the bubbling experiment was the same as above, and the obtained particles were free-dried (−80 °C for 24 h) for further characterization.

3. Results and Discussion

3.1. Removal of Target Pollutants in the Washing Solution of FA

Figure 1 presents the XRD patterns, SEM images, EDS analysis, and X-ray mapping images before and after the washing pre-treatment for FA. As shown in Figure 1a, the crystalline substances in the original FA could be divided into two categories, mainly including alkali chlorides (e.g., NaCl, KCl, and CaClOH) and Ca-containing phases (e.g., CaCO3, Ca(OH)2, CaSO4, and CaClOH [36]). Based on the XRD peaks shown in Figure 1a, we infer that CaCO3 could be ascribed to the natural carbonation during the storage of FA. The presences of CaSO4 and CaClOH were mainly due to the adsorption of acid gases (e.g., HCl and SO3) in the exhaust gas by adding lime (Equations (1) and (2)) [37]. Moreover, the residual alkaline compounds, including CaSO4, CaCO3, Ca(OH)2, and CaClOH [13], could be further applied to sequestrate CO2 via a carbonation reaction, indicating the potential of FA for capturing CO2. Furthermore, the composition of the original FA differed greatly compared to the washed FA. Therefore, we conclude that most soluble salts, including NaCl and KCl [38], could be removed by washing. After the water washing process, Ca(OH)2 were gradually generated according to Equations (3)–(5). Based on the high XRD intensity of Ca(OH)2 shown in Figure 1a, we preferred to optimize L/S to 8 mL/g for the following characterization. For comparison, the reaction in Equation (3) led to the complete disappearance of CaClOH in XRD.
CaO + HCl = CaClOH
CaO(S) + SO3 = CaSO4
2CaClOH + nH2O = Ca(OH)2(S) + CaCl2∙nH2O
Ca3Si2O7 + 7H2O = 3Ca(OH)2(S) + 2H4SiO4
CaO(S) + H2O = Ca(OH)2(S)
Figure 1b,c reveal the SEM variations of the original and water-washed FA. A spherical and layer-shaped structure [39] of the original FA was observed with irregular shapes compared with a dense and square structure. This indicates a significant change in the microstructures of FA after pre-treatment with washing. Moreover, the EDS spectra and X-ray mapping images shown in Figure 1d–g confirm that a washing pre-treatment could achieve the effective removal of most chlorine and soluble salts (e.g., Cl, Na+, and K+), accompanied by the leach of HMs. Thus, the EDS spectral variations indicate that the residual concentrations of Pb, Cu, Zn, and Cd in the washed FA apparently decreased compared with the original FA. Furthermore, Figure 1h,i reveal the variations of leaching concentrations of target HMs and pH values in the washing solution. Based on the results of ICP-MS shown in the inset of Figure 1h, we conclude that fly ash had an extremely high Zn content of 5278.5 mg/kg, which is in agreement with the available document [40]. The contents of Pb, Cu, Cr, and Cd could reach up to 2251.4 mg/kg, 1426.5 mg/kg, 103.2 mg/kg, and 96.9 mg/kg, respectively. Notably, the leaching concentrations of target HMs are slightly influenced by the temperature of the washing solution. As shown in Figure 1h, the maximal leaching concentration of target HMs in washing solution at 25 °C could reach up to 15.8 mg/L of Pb, compared with 0.64 mg/L of Cu, 0.38 mg/L of Zn, 1.37 mg/L of Cr, and 0.24 mg/L of As. This generally agrees with the prior study [41]. Furthermore, the terminal pH values of the washing solution shown in Figure 1i are basically within the range of 11.4–13.3. This indicates that the washing solution of the original FA, including Pb concentrations and pH values, did not meet the integrated wastewater discharge standard in GB 8978−1996 [42]. Thus, this emphasizes the necessity for controlling the micro-pollutions, mainly including Pb, in the washing solution of FA.
Figure 2a−c show the removal efficiency for Pb ions and variations of pH values versus the main parameters of bubbling simulated gases into the washing solution of FA. The results show that the pH values of the washing solution slowly decreased in no more than 30 min of bubbling time, and a further increase in the bubbling time could lead to a clear downward trend in pH values. This is mainly ascribed to the high initial pH values in the washing solution of FA (approx. 13.3), which could hardly be neutralized by bubbling gases. For comparison, the carbonation reaction of 40 min could decrease the pH values of the washing solution to approximately 7. Furthermore, the effects of gas flow rates, temperatures, and different CO2 ratios on the removal efficiency of Pb ions are also recorded. As shown in Figure 2a,b, the removal efficiency of Pb ions increased from 82.2% to 97.9% when the flow rates increased from 10 to 15 sccm within 30 min of the gas bubbling process at 25 °C. In contrast, the same gas flow rates conducted at 50 °C and 80 °C had a negative influence on the removal of Pb. As for temperature, the results shown in Figure 2a–c verify that the removal efficiency of Pb significantly decreased as the carbonation temperature increased from 25 °C to 80 °C. For example, the maximum removal efficiency of Pb was up to 97.9% at 25 °C after 29 min of bubbling. On the contrary, the removal efficiency of Pb decreased to 78.1% at 80 °C, accompanied by an increase in the bubbling time by at least 10 min (flow rate = 15 sccm, CO2 ratio = 33 mol%). The above results also indicate that a higher temperature is not favorable for Pb removal in the washing solution of FA. The high temperature is detrimental to the dissolution of CO32−, which is regarded as a rate-limiting step in the carbonation process of the washing solution. Therefore, temperature could control the solubility of H2CO3 [43] and further determine the carbonation for Pb removal. Additionally, Figure 2d−f present the variations of Pb versus pH values. The results shown in Figure 2d−f verify that the concentration of Pb reduced sharply within a pH of 10.25 and then decreased slowly at a temperature of 25 °C. This was consistent with the trend at other temperatures (approx. 80 °C) under the condition of 10 sccm−33 mol%. Similar phenomena could be observed in other experimental parameters (15 sccm−33 mol%, 15 sccm−15 mol%). This could be because the species distributions of H2CO3 could be directly determined by the pH values of the washing solution. Consequently, the preferred flow rate of gas and temperature were optimized as 15 sccm and 25 °C, respectively.
Under the preferred condition of 25 °C and 15 sccm gas flow rate, two simulated flue gases could significantly remove 97.9−99.2% of Pb (in Figure 2b,c). Furthermore, the maximum removal efficiency of Zn (99%) and Cu (81%) could also be observed under the same conditions (Figure S2). However, the removal efficiency of Cr and As showed no obvious patterns (Figure S2). After the bubbling treatment, the residual Pb concentration of 0.34 mg/L, as well as terminal pH in the washing solution of FA, met the regulatory standard of integrated wastewater discharge standard in GB 8978−1996 [42] (e.g., Pb: 1.0 mg/L, pH: 6−8). Therefore, it is feasible to use flue gas emitted from the CFPP and cement/steel industries to remediate the micro-pollutions in the washing solution of FA.

3.2. Kinetics Study Analysis

Figure 3a presents the variations of Pb ions concentrations versus the contact time. The results in Figure 3a reveal that within the initial 40 min of reaction, the removal of Pb ions basically fit with a pseudo-first-order kinetics model described in Equation (6). Furthermore, the maximal kinetic constant shown in the inset of Figure 3a was calculated as 1.31 × 10−1 min−1 under the preferred conditions of 25 °C, and 15 sccm−15 mol% CO2. Moreover, the kinetic constants of Pb ions decreased with increasing reaction temperature. For example, the maximal kinetic constants of Pb ions were classified as follows: 1.3 × 10−1 min−1 (at 25 °C) > 4.1 × 10−2 min−1 (at 50 °C) > 3.1 × 10−2 min−1 (at 80 °C). This has also coincided with the results shown in Figure 2d–f, which presents Pb ions’ variations versus pH values.
ln ( C t C 0 ) = k o b s × t
where Ct and C0 are the concentrations of Pb (mg/L) after a reaction time of t and 0 (min), respectively; k is the rate constant (min−1) of the pseudo-first-order kinetics equation, and t is the reaction time (min).
To explore the thermodynamic parameters for the carbonation process, free energy change (ΔG0), enthalpy (ΔH0), and entropy change (ΔS0) were calculated by the van ’t Hoff equations shown in Equations (7)–(9) [44].
ln K D = Δ S 0 R Δ H 0 R T
K D = C 0 C e C e
Δ G 0 = Δ H 0 T Δ S 0
where R is the universal gas constant (8.314 J mol−1 K−1), T is the absolute temperature (K), and KD (L/mol) is the distribution coefficient at different temperature levels. KD (mL/g), ΔS0 [J/(mol K)], and ΔH0 (kJ/mol) represent the reaction coefficient, the standard entropy, and the standard enthalpy, respectively. C0 and Ct (mg/L) are the concentrations of Pb at times 0 and t. The change in enthalpy (ΔH0 in kJ/mol) and entropy (ΔS0 in J/molK) could be calculated from linear coefficients using van ’t Hoff’s equation.
Figure 3b shows the plot of ΔG0 versus reaction temperature within 298.15–353.15 K. As shown in Table 1, we could determine that for removing Pb ions, ΔG0, ΔH0, and ΔS0 are negative, negative, and negative, respectively. This indicates that the removal of Pb ions was a spontaneous, exothermic, and entropy-decreasing process. The absolute ΔG0 values increased with increasing reaction temperature. This could be because of bubbling CO2 into the FA washing solution and the precipitation of deposits of lead and calcium carbonates. This could apparently decrease the level of disorder and randomness at the gas/solution interface. More exploration will be conducted in the following section.

3.3. Carbonation Mechanism

To elucidate the Pb precipitation behaviors in the washing solution system during the carbonation process (e.g., Pb2+ and CO32−), we adopted the Visual MINTEQ (ver.3) to conduct the equilibrium calculations and examine the potential speciation distribution and saturation indices (SIs) versus different pH values [45]. Figure 4a–c reveal the main speciation distribution of Pb ions during the carbonation process. As shown in Figure 4a–c, the removal of Pb ions was due to the lower pH of the solution induced by bubbling CO2, accompanied by a series of changes in the chemistry and mineralogy [33]. The carbonation process occurred mainly via two consecutive steps: the solubility of Pb ions against low solution pH [46] and the transformation of Pb hydroxides into Pb carbonates. For example, Pb(OH)3 ions were significant species of Pb ions in the washing solution under strongly alkaline conditions (e.g., pH > 12). As the dissolution of CO2 led to a decrease in pH [47] (e.g., pH decreased from 13.3 to 9.5), Pb(OH)3 ions gradually transformed into Pb(OH)2(aq) and Pb(OH). The aforementioned process was accompanied by the increases of Pb2+ and CO32− ions in the washing solution (2Pb(OH)3 + H2CO3(aq) = 2Pb(OH)2(aq) + CO32−; 2Pb(OH)2 + H2CO3(aq) = 2Pb(OH)+ + CO32-; 2Pb(OH)+ + H2CO3(aq) = 2Pb2+ + CO32−). Furthermore, Pb hydroxides could be further transformed into PbCO3 due to excessive carbonate ions (Pb(OH)2(aq) + H2CO3(aq) = PbCO3(S) + 2H2O; Pb2+ + CO32− = PbCO3(S)). Furthermore, the EDS images shown in Figure 4d–f verify that the components of white precipitants mainly included Ca element (e.g., 93.5−96.3 wt%) as well as Pb element (e.g., 1.9−3.3 wt%), in addition to the slight residual amounts of Al, Fe, As, and Zn elements. Furthermore, the variations of the SEM images shown in the insets of Figure 4d–f verify that the rectangular particles of precipitants gradually transformed into circular particles as the bubbling rates of CO2 increased. Based on the classical nucleation and growth theory, we could conclude that CaCO3 particles originate with the nucleation in a supersaturated solution and further grow into microcrystals [48]. For example, the relatively high nucleation rate could result in a small size of the obtained CaCO3 shown in Equations (10) and (11) [49] and the supersaturation ratio (S) in the interface is expressed as in Equation (12).
r = 2 γ v K T ln S  
J = A   e x p [   - 16 γ 3 v 2 π   / 3   K 3 T 2 ( ln S   ) 2   ]  
S = α C O 3 2 + α C a 2 + K s p
where r is the radius of the critical nucleus, γ is interfacial free energy, ν is the volume of a molecule inside the nucleus, K and T are Boltzmann constant and absolute temperature, and J is the nucleation rate, respectively; Ksp is the thermodynamic constant of CaCO3 and α is ionic activity coefficient [50].
Figure 4g,h compare the XPS spectral variations of precipitants. The XPS spectra show that binding energies (BEs) at 138.45 and 143.38 eV for Pb 4f doublet peaks corresponded to PbCO3 [51]. Furthermore, the BE at 144.78 and 139.68 eV were attributed to Pb3(CO3)2(OH)2 [52]. We infer that the co-precipitation of PbCO3 and Pb(OH)2 at high alkalinity led to the generation of the aforementioned Pb3(CO3)2(OH)2. Furthermore, the XPS spectra in Figure 4g,h coincided with the EDS images in Figure 4d–f, verifying the co-existence of Pb and calcium on the surface of precipitants. Furthermore, PbCO3 was mainly generated with the CO2 bubbling process, disrupting the equilibrium of the Pb complex in the alkaline washing solutions of FA. A series of carbonate precipitation processes have been proposed in Figure 4i. This indicates that the washing solution of FA has a significant capacity for CO2 capture, making it a promising candidate for efficient CO2 capture.
Table 2 summarizes the main parameters (e.g., alkaline waste source, HMs solidification, carbonation efficiency, and obtained CaCO3) in available carbonation processes documented with alkaline residues as feedstocks. Based on the comparisons shown in Table 2, we ascribe the different variations of carbonated capacity for alkaline solid waste to different residue CaO contents in the alkaline solid waste. In addition, the state-of-the-art carbon capture, utilization, and storage (CCUS) techniques focus on mineral carbonation mainly via gas–solid [8,13,53] or gas–liquid–solid [36,37,54,55,56,57,58] processes. To our best knowledge, gas–liquid–solid reactions of CO2 with fly ash could achieve rapid transient mass transfer compared with gas–solid reactions under high pressure and high-temperature conditions [27]. However, the newly adopted thermal detoxification of dioxin in FA, including a microwave-initiated thermal process [29], could destroy the mineral carbonation and lead to the release of CO2. For comparison, bubbling the pure CO2 or simulated fuel gas in effluent could theoretically remove the residual HMs, and neutralize the alkalinity [28,41,59] in an aqueous solution. Meanwhile, the obtained mineral carbonate generated from bubbling the pure CO2 could also be utilized for further purification of wastewater. Currently, indirect carbonation is an important process to achieve CO2 storage and recover CaCO3. It focuses on using various leaching agents, such as acids [21], alkalis, and salts [24,25,26,60], to maximize the extraction of calcium components from the alkaline waste and produce high purity CaCO3 with specific shapes.

3.4. Characteristics of Obtained Particle

Figure 5a presents the XRD patterns of obtained particles. By comparing with the aforementioned two reference patterns of CaCO3, we infer that the obtained particles were mainly composed of calcite (JCPDS No. 47−1743) and vaterite (JCPDS No. 33−0268) microcrystalline. The SEM images presented in Figure 5b,c further show that the obtained particles were aggregates of smaller units. With increasing flow rates of CO2, the ratios of vaterite to calcite gradually increased, but suddenly dropped at the highest flow rate (Figure 5d). Previously, the polymorph formation of CaCO3 has been reported to be highly influenced by the precipitation conditions, including pH, supersaturation, temperature, and additives [61]. In this study, different CO2 flow rates may affect both pH and supersaturation in the solution. Thus, the transformation between vaterite and calcite of CaCO3 could be ascribed to a result of compound effects [61,62]. More importantly, vaterite has been reported to excel in removing toxic metal ions, including Pb2+ [63,64]. Correspondingly, the vaterite contents were higher under 15 sccm flow rate of CO2 than 10 sccm (Figure 5b). As shown in Figure 2a,b, we found that under the simulated flue gas at 15 sccm, the removal efficiency of Pb ions was better compared with that at 10 sccm. In the future, the gas flow rates of 30 sccm and 45 sccm should also be considered for the HMs removal in the washing solution.
The XRD patterns and SEM images indicate that the obtained particles were mainly composed of vaterite and calcite of CaCO3. The EDS images in Figure 4d–f and the XPS spectra in Figure 4h verifies the existence of a small amount of Pb in these obtained particles. Therefore, the toxicity of obtained particles needed to be evaluated, and the associated experiments were described in the Supplementary Materials. The acute toxicity experiment demonstrates that the mortality rates of zebrafish were 0% after exposure to obtained particles for 24 h, 48 h, 72 h, and 96 h, respectively. Given that zebrafish is an excellent model animal for human health research [65], we infer that the obtained particles from the washing solution are safe for further application. Moreover, we used 50 mg of obtained particles for the adsorption of 50 mL Congo red dye (initial concentration: 10 mg/L, 15 mg/L, 20 mg/L, and 30 mg/L, respectively), which was conducted in a thermostatic oscillator (e.g., 150 rpm, 24 h). The results shown in Figure S3 verify that obtained CaCO3 particles could function as promising adsorbents.

4. Conclusions

The simulated exhaust gases of CFPP and cement/steel industry could achieve the remediation of micro-pollution, including Pb and alkalinity in the washing solution for MSWI fly ash (e.g., Pb: 14.8 mg/L-15.8 mg/L; solution pH: 12.8–13.6). After the bubbling treatment with simulated exhaust gases in the washing solution, the residual concentration of HMs, including Pb ions and terminal pH, basically met the regulatory threshold values of integrated wastewater discharge standard in China (Pb: 1.0 mg/L, pH = 6–9). The removal of Pb ions was a spontaneous, exothermic, and entropy-decreasing process, and the proposed removal mechanism was ascribed to the generation of PbCO3 and Pb3(OH)2(CO3)2. The obtained CaCO3 was proved non-toxic by acute toxicity test and showed a quick adsorption performance of Congo red dye. This study demonstrates the possibility of simulated CO2 exhaust gas to remediate micro-pollution of FA washing solution, and provides a theoretical basis for the reuse of real exhaust gas containing CO2. In the context of carbon neutrality and emission peak, the effective utilization of exhaust gas containing CO2 can reduce the HM pollution of FA washing solution, sequestrate CO2, and recover CaCO3.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/su15075873/s1, Figure S1: Schematic process flow for MSWI fly ash (FA) carbonation; Figure S2: (a–c) Variations in Cu removal efficiency versus bubbling time of simulated exhaust gas; (d–f) variations in Zn removal efficiency versus bubbling time of simulated exhaust gas; (g–i) variations in As removal efficiency versus bubbling time of simulated exhaust gas; (j–l) variations in Cr removal efficiency versus bubbling time of simulated exhaust gas. Note: the preferred ratio between liquid and solid, L/S = 8 mL/g, washing time = 10 min, washing temperature = 25 °C, 50 °C and 80 °C, respectively. The CO2 proportions in simulated exhaust gases were 15 mol% and 33 mol%, and flow rates were selected as 10 sccm and 15 sccm, respectively; Figure S3: Removal of Congo red dye by particles obtained under CO2 aeration with different flow rates (10, 15, 30, 45, and 61 sccm).

Author Contributions

Conceptualization, Y.J.; Methodology, L.W.; Validation, Y.T.; Formal analysis, Y.G.; Investigation, Y.T.; Resources, Y.J.; Data curation, Y.T.; Writing—original draft preparation, Y.T.; Funding acquisition, Y.J.; Visualization, Y.T.; Writing—review and editing, X.L., W.X., Y.Z., Y.G. and Y.J.; Supervision, Y.J., X.S., D.K., Y.G. and L.S. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the fundings provided by the National Key Research and Development Program of China (Grant No.: 2019YFE0111100), the Central Scientific Research Projects for Public Welfare Research Institutes (Grant No.: GYZX210301), and the National Natural Science Foundation of China (Grant No.: 52170137).

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We sincerely thank the editors and the four anonymous reviewers for their constructive suggestions and comments.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Wang, L.; Zhang, Y.; Chen, L.; Guo, B.; Tan, Y.; Sasaki, K.; Tsang, D.C.W. Designing Novel Magnesium Oxysulfate Cement for Stabilization/Solidification of Municipal Solid Waste Incineration Fly Ash. J. Hazard. Mater. 2022, 423, 127025. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, Z.; Zhang, S.; Lin, X.; Li, X. Decomposition and Reformation Pathways of PCDD/Fs during Thermal Treatment of Municipal Solid Waste Incineration Fly Ash. J. Hazard. Mater. 2020, 394, 122526. [Google Scholar] [CrossRef] [PubMed]
  3. Le, N.H.; Razakamanantsoa, A.; Nguyen, M.L.; Phan, V.T.; Dao, P.L.; Nguyen, D.H. Evaluation of Physicochemical and Hydromechanical Properties of MSWI Bottom Ash for Road Construction. Waste Manag. 2018, 80, 168–174. [Google Scholar] [CrossRef] [PubMed]
  4. Qin, J.; Zhang, Y.; Yi, Y.; Fang, M. Carbonation of Municipal Solid Waste Gasification Fly Ash: Effects of Pre-Washing and Treatment Period on Carbon Capture and Heavy Metal Immobilization. Environ. Pollut. 2022, 308, 119662. [Google Scholar] [CrossRef]
  5. Jozewicz, W.; Chang, J.C.S.; Sedman, C.B. Bench–Scale Evaluation of Calcium Sorbents for Acid Gas Emission Control. Environ. Prog. 1990, 9, 137–142. [Google Scholar] [CrossRef]
  6. Zhang, T.A.; Zhao, H.; Liu, Y.; Dou, Z.; Lv, G.; Zhao, Q.; Li, Y. Recent Advances in Carbon Dioxide Mineralization to Nano-Size Calcium Carbonate Utilizing Wastewater. In Energy Technology 2014; Wang, C.W., de Bakker, J., Belt, C.K., Jha, A., Neelameggham, N.R., Pati, S., Prentice, L.H., Tranell, G., Brinkman, K.S., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2014; pp. 103–110. [Google Scholar] [CrossRef]
  7. IEA. Global Energy Review: CO2 Emissions in 2021—Analysis. Available online: https://www.iea.org/reports/global-energy-review-co2-emissions-in-2021-2 (accessed on 18 December 2022).
  8. Gu, Q.; Wang, T.; Wu, W.; Wang, D.; Jin, B. Influence of Pretreatments on Accelerated Dry Carbonation of MSWI Fly Ash under Medium Temperatures. Chem. Eng. J. 2021, 414, 128756. [Google Scholar] [CrossRef]
  9. Zdeb, J.; Howaniec, N.; Smolinski, A. Experimental Study on Combined Valorization of Bituminous Coal Derived Fluidized Bed Fly Ash and Carbon Dioxide from Energy Sector. Energy 2023, 265, 126367. [Google Scholar] [CrossRef]
  10. Yuan, Q.; Zhang, Y.; Wang, T.; Wang, J. Characterization of Heavy Metals in Fly Ash Stabilized by Carbonation with Supercritical CO2 Coupling Mechanical Force. J. CO2 Util. 2023, 67, 102308. [Google Scholar] [CrossRef]
  11. Abe, M.; Tanaka, S.; Noguchi, M.; Yamasaki, A. Investigation of Mineral Carbonation with Direct Bubbling into Concrete Sludge. ACS Omega 2021, 6, 15564–15571. [Google Scholar] [CrossRef]
  12. Liu, W.; Teng, L.; Rohani, S.; Qin, Z.; Zhao, B.; Xu, C.C.; Ren, S.; Liu, Q.; Liang, B. CO2 Mineral Carbonation Using Industrial Solid Wastes: A Review of Recent Developments. Chem. Eng. J. 2021, 416, 129093. [Google Scholar] [CrossRef]
  13. Tamilselvi Dananjayan, R.R.; Kandasamy, P.; Andimuthu, R. Direct Mineral Carbonation of Coal Fly Ash for CO2 Sequestration. J. Clean. Prod. 2016, 112, 4173–4182. [Google Scholar] [CrossRef]
  14. 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]
  15. Chen, W.; Wang, Y.; Sun, Y.; Fang, G.; Li, Y. Release of Soluble Ions and Heavy Metal during Fly Ash Washing by Deionized Water and Sodium Carbonate Solution. Chemosphere 2022, 307, 135860. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, J.; Shen, Y.; Chen, Z.; Fu, C.; Li, M.; Mao, T.; Xu, R.; Lin, X.; Li, X.; Yan, J. Accelerated Carbonation of Ball-Milling Modified MSWI Fly Ash: Migration and Stabilization of Heavy Metals. J. Environ. Chem. Eng. 2023, 11, 109396. [Google Scholar] [CrossRef]
  17. 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]
  18. Tian, S.; Jiang, J. Sequestration of Flue Gas CO2 by Direct Gas–Solid Carbonation of Air Pollution Control System Residues. Environ. Sci. Technol. 2012, 46, 13545–13551. [Google Scholar] [CrossRef]
  19. Zhang, 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]
  20. Eloneva, S.; Teir, S.; Salminen, J.; Fogelholm, C.J.; Zevenhoven, R. Fixation of CO2 by Carbonating Calcium Derived from Blast Furnace Slag. Energy 2008, 33, 1461–1467. [Google Scholar] [CrossRef]
  21. 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]
  22. Kunzler, C.; Alves, N.; Pereira, E.; Nienczewski, J.; Ligabue, R.; Einloft, S.; Dullius, J. CO2 Storage with Indirect Carbonation Using Industrial Waste. Energy Procedia 2011, 4, 1010–1017. [Google Scholar] [CrossRef] [Green Version]
  23. Kang, J.M.; Murnandari, A.; Youn, M.H.; Lee, W.; Park, K.T.; Kim, Y.E.; Kim, H.J.; Kang, S.P.; Lee, J.H.; Jeong, S.K. Energy-Efficient Chemical Regeneration of AMP Using Calcium Hydroxide for Operating Carbon Dioxide Capture Process. Chem. Eng. J. 2018, 335, 338–344. [Google Scholar] [CrossRef]
  24. Li, W.; Huang, Y.; Wang, T.; Fang, M.; Li, Y. Preparation of Calcium Carbonate Nanoparticles from Waste Carbide Slag Based on CO2 Mineralization. J. Clean. Prod. 2022, 363, 132463. [Google Scholar] [CrossRef]
  25. Kodama, S.; Nishimoto, T.; Yamamoto, N.; Yogo, K.; Yamada, K. Development of a New PH-Swing CO2 Mineralization Process with a Recyclable Reaction Solution. Energy 2008, 33, 776–784. [Google Scholar] [CrossRef]
  26. Jeon, J.; Kim, M.J. CO2 Storage and CaCO3 Production Using Seawater and an Alkali Industrial By-Product. Chem. Eng. J. 2019, 378, 122180. [Google Scholar] [CrossRef]
  27. Qin, J.; Zhang, Y.; Yi, Y.; Fang, M. Carbonation Treatment of Gasification Fly Ash from Municipal Solid Waste Using Sodium Carbonate and Sodium Bicarbonate Solutions. Environ. Pollut. 2022, 299, 118906. [Google Scholar] [CrossRef] [PubMed]
  28. Kang, D.; Son, J.; Yoo, Y.; Park, S.; Huh, I.S.; Park, J. Heavy-Metal Reduction and Solidification in Municipal Solid Waste Incineration (MSWI) Fly Ash Using Water, NaOH, KOH, and NH4OH in Combination with CO2 Uptake Procedure. Chem. Eng. J. 2020, 380, 122534. [Google Scholar] [CrossRef]
  29. Ju, Y.; Zhang, K.; Yang, T.; Deng, D.; Qiao, J.; Wang, P.; You, Y.; Du, L.; Chen, G.; Kołodyńska, D.; et al. The Influence of a Washing Pretreatment Containing Phosphate Anions on Single-Mode Microwave-Based Detoxification of Fly Ash from Municipal Solid Waste Incinerators. Chem. Eng. J. 2020, 387, 124053. [Google Scholar] [CrossRef]
  30. Deng, D.; Qiao, J.; Liu, M.; Kołodyńska, D.; Zhang, M.; Dionysiou, D.D.; Ju, Y.; Ma, J.; Chang, M.-b. Detoxification of Municipal Solid Waste Incinerator (MSWI) Fly Ash by Single-Mode Microwave (MW) Irradiation: Addition of Urea on the Degradation of Dioxin and Mechanism. J. Hazard. Mater. 2019, 369, 279–289. [Google Scholar] [CrossRef]
  31. Ju, Y.; Deng, D.; Li, H.; You, Y.; Qiao, J.; Su, L.; Xing, W.; Liang, M.; Dionysiou, D.D. Rapid Detoxification of Dioxin and Simultaneous Stabilization of Targeted Heavy Metals: New Insight into a Microwave-Induced Pyrolysis of Fly Ash. Chem. Eng. J. 2022, 429, 131939. [Google Scholar] [CrossRef]
  32. Chen, Q.; Luo, Z.; Hills, C.; Xue, G.; Tyrer, M. Precipitation of Heavy Metals from Wastewater Using Simulated Flue Gas: Sequent Additions of Fly Ash, Lime and Carbon Dioxide. Water Res. 2009, 43, 2605–2614. [Google Scholar] [CrossRef]
  33. Wang, L.; Chen, Q.; Jamro, I.A.; Lia, R.D.; Baloch, H.A. Accelerated Co-Precipitation of Lead, Zinc and Copper by Carbon Dioxide Bubbling in Alkaline Municipal Solid Waste Incinerator (MSWI) Fly Ash Wash Water. RSC Adv. 2016, 6, 20173–20186. [Google Scholar] [CrossRef]
  34. Dal Pozzo, A.; Armutlulu, A.; Rekhtina, M.; Müller, C.R.; Cozzani, V. CO2 Uptake Potential of Ca-Based Air Pollution Control Residues over Repeated Carbonation–Calcination Cycles. Energy Fuels 2018, 32, 5386–5395. [Google Scholar] [CrossRef]
  35. Han, S.J.; Im, H.J.; Wee, J.H. Leaching and Indirect Mineral Carbonation Performance of Coal Fly Ash-Water Solution System. Appl. Energy 2015, 142, 274–282. [Google Scholar] [CrossRef]
  36. Ni, P.; Xiong, Z.; Tian, C.; Li, H.; Zhao, Y.; Zhang, J.; Zheng, C. Influence of Carbonation under Oxy-Fuel Combustion Flue Gas on the Leachability of Heavy Metals in MSWI Fly Ash. Waste Manag. 2017, 67, 171–180. [Google Scholar] [CrossRef]
  37. Zhu, Z.; Guo, Y.; Zhao, Y.; Zhou, T. Carbon Reclamation from Biogas Plant Flue Gas for Immobilizing Lead and Neutralizing Alkalis in Municipal Solid Waste Incineration Fly Ash. Chem. Eng. J. 2022, 435, 134812. [Google Scholar] [CrossRef]
  38. Zheng, R.; Wang, Y.; Liu, Z.; Zhou, J.; Yue, Y.; Qian, G. Environmental and Economic Performances of Municipal Solid Waste Incineration Fly Ash Low-Temperature Utilization: An Integrated Hybrid Life Cycle Assessment. J. Clean. Prod. 2022, 340, 130680. [Google Scholar] [CrossRef]
  39. Li, H.; Sun, J.; Gui, H.; Xia, D.; Wang, Y. Physiochemical Properties, Heavy Metal Leaching Characteristics and Reutilization Evaluations of Solid Ashes from Municipal Solid Waste Incinerator Plants. Waste Manag. 2022, 138, 49–58. [Google Scholar] [CrossRef]
  40. Li, Y.; Zhang, J.; Liu, Z.; Chen, L.; Wang, Y. Harmless Treatment of Municipal Solid Waste Incinerator Fly Ash through Shaft Furnace. Waste Manag. 2021, 124, 110–117. [Google Scholar] [CrossRef]
  41. Yang, Z.; Tian, S.; Ji, R.; Liu, L.; Wang, X.; Zhang, Z. Effect of Water-Washing on the Co-Removal of Chlorine and Heavy Metals in Air Pollution Control Residue from MSW Incineration. Waste Manag. 2017, 68, 221–231. [Google Scholar] [CrossRef]
  42. GB8978-1996; Integrated Wastewater Discharge Standard. State Environmental Protection Agency: Beijing, China, 1996.
  43. Carroll, J.J.; Slupsky, J.D.; Mather, A.E. The Solubility of Carbon Dioxide in Water at Low Pressure. J. Phys. Chem. Ref. Data 1991, 20, 1201–1209. [Google Scholar] [CrossRef]
  44. Zhang, Y.; Lin, S.; Qiao, J.; Kołodyńska, D.; Ju, Y.; Zhang, M.; Cai, M.; Deng, D.; Dionysiou, D.D. Malic Acid-Enhanced Chitosan Hydrogel Beads (MCHBs) for the Removal of Cr(VI) and Cu(II) from Aqueous Solution. Chem. Eng. J. 2018, 353, 225–236. [Google Scholar] [CrossRef]
  45. Zhou, N.; Luther, G.W.; Chan, C.S. Ligand Effects on Biotic and Abiotic Fe(II) Oxidation by the Microaerophile Sideroxydans Lithotrophicus. Environ. Sci. Technol. 2021, 55, 9362–9371. [Google Scholar] [CrossRef] [PubMed]
  46. Wu, K.J.; Tse, E.C.M.; Shang, C.; Guo, Z. Nucleation and Growth in Solution Synthesis of Nanostructures—From Fundamentals to Advanced Applications. Prog. Mater. Sci. 2022, 123, 100821. [Google Scholar] [CrossRef]
  47. Van Gerven, T.; Moors, J.; Dutré, V.; Vandecasteele, C. Effect of CO2 on Leaching from a Cement-Stabilized MSWI Fly Ash. Cem. Concr. Res. 2004, 34, 1103–1109. [Google Scholar] [CrossRef]
  48. Trushina, D.B.; Bukreeva, T.V.; Kovalchuk, M.V.; Antipina, M.N. CaCO3 Vaterite Microparticles for Biomedical and Personal Care Applications. Mater. Sci. Eng. C 2014, 45, 644–658. [Google Scholar] [CrossRef] [PubMed]
  49. Trushina, D.B.; Bukreeva, T.V.; Antipina, M.N. Size-Controlled Synthesis of Vaterite Calcium Carbonate by the Mixing Method: Aiming for Nanosized Particles. Cryst. Growth Des. 2016, 16, 1311–1319. [Google Scholar] [CrossRef]
  50. Fernandez-Martinez, A.; Hu, Y.; Lee, B.; Jun, Y.S.; Waychunas, G.A. In Situ Determination of Interfacial Energies between Heterogeneously Nucleated CaCO3 and Quartz Substrates: Thermodynamics of CO2 Mineral Trapping. Environ. Sci. Technol. 2013, 47, 102–109. [Google Scholar] [CrossRef]
  51. Jia, K.; Feng, Q.; Zhang, G.; Ji, W.; Zhang, W.; Yang, B. The Role of S(II) and Pb(II) in Xanthate Flotation of Smithsonite: Surface Properties and Mechanism. Appl. Surf. Sci. 2018, 442, 92–100. [Google Scholar] [CrossRef]
  52. Zhang, S.; Wen, S.; Xian, Y.; Zhao, L.; Feng, Q.; Bai, S.; Han, G.; Lang, J. Lead Ion Modification and Its Enhancement for Xanthate Adsorption on Smithsonite Surface. Appl. Surf. Sci. 2019, 498, 143801. [Google Scholar] [CrossRef]
  53. Chiang, K.Y.; Hu, Y.H. Water Washing Effects on Metals Emission Reduction during Municipal Solid Waste Incinerator (MSWI) Fly Ash Melting Process. Waste Manag. 2010, 30, 831–838. [Google Scholar] [CrossRef]
  54. Jiang, J.; Chen, M.; Zhang, Y.; Xu, X. Pb Stabilization in Fresh Fly Ash from Municipal Solid Waste Incinerator Using Accelerated Carbonation Technology. J. Hazard. Mater. 2009, 161, 1046–1051. [Google Scholar] [CrossRef] [PubMed]
  55. Ukwattage, N.L.; Ranjith, P.G.; 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]
  56. Liu, W.; Su, S.; Xu, K.; Chen, Q.; Xu, J.; Sun, Z.; Wang, Y.; Hu, S.; Wang, X.; Xue, Y.; et al. CO2 Sequestration by Direct Gas–Solid Carbonation of Fly Ash with Steam Addition. J. Clean. Prod. 2018, 178, 98–107. [Google Scholar] [CrossRef]
  57. 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]
  58. Chen, J.; Lin, X.; Li, M.; Mao, T.; Li, X.; Yan, J. Heavy Metal Solidification and CO2 Sequestration from MSWI Fly Ash by Calcium Carbonate Oligomer Regulation. J. Clean. Prod. 2022, 359, 132044. [Google Scholar] [CrossRef]
  59. Wang, L.; Jin, Y.; Nie, Y.; Li, R. Recycling of Municipal Solid Waste Incineration Fly Ash for Ordinary Portland Cement Production: A Real-Scale Test. Resour. Conserv. Recycl. 2010, 54, 1428–1435. [Google Scholar] [CrossRef]
  60. Tong, Z.; Ma, G.; Zhou, D.; Yang, G.; Peng, C. The Indirect Mineral Carbonation of Electric Arc Furnace Slag Under Microwave Irradiation. Sci. Rep. 2019, 9, 7676. [Google Scholar] [CrossRef] [Green Version]
  61. Zhou, G.T.; Yao, Q.Z.; Fu, S.Q.; Guan, Y.B. Controlled Crystallization of Unstable Vaterite with Distinct Morphologies and Their Polymorphic Transition to Stable Calcite. Eur. J. Mineral. 2010, 22, 259–269. [Google Scholar] [CrossRef] [Green Version]
  62. Han, Y.S.; Hadiko, G.; Fuji, M.; Takahashi, M. Factors Affecting the Phase and Morphology of CaCO3 Prepared by a Bubbling Method. J. Eur. Ceram. Soc. 2006, 26, 843–847. [Google Scholar] [CrossRef]
  63. Dang, H.C.; Yuan, X.; Xiao, Q.; Xiao, W.X.; Luo, Y.K.; Wang, X.L.; Song, F.; Wang, Y.Z. Facile Batch Synthesis of Porous Vaterite Microspheres for High Efficient and Fast Removal of Toxic Heavy Metal Ions. J. Environ. Chem. Eng. 2017, 5, 4505–4515. [Google Scholar] [CrossRef]
  64. Lin, P.Y.; Wu, H.M.; Hsieh, S.L.; Li, J.S.; Dong, C.; Chen, C.W.; Hsieh, S. Preparation of Vaterite Calcium Carbonate Granules from Discarded Oyster Shells as an Adsorbent for Heavy Metal Ions Removal. Chemosphere 2020, 254, 126903. [Google Scholar] [CrossRef] [PubMed]
  65. Liu, W.; Huang, G.; Su, X.; Li, S.; Wang, Q.; Zhao, Y.; Liu, Y.; Luo, J.; Li, Y.; Li, C.; et al. Zebrafish: A Promising Model for Evaluating the Toxicity of Carbon Dot-Based Nanomaterials. ACS Appl. Mater. Interfaces 2020, 12, 49012–49020. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) The XRD patterns of the fly ash treated with different L/S; (b,c) SEM images of the original and washed fly ash; (dg) the X-ray mapping of the original fly ash and washed fly ash (with the main element contents shown in the inset); (h) variations in leaching concentrations of target HMs; (i) final solution pH values conducted at different temperatures (with the trace elements of the original fly ash shown in the inset of (h)). Note: L/S = 8 mL/g, washing time = 10 min, washing temperature = 25 °C, 50 °C, and 80 °C, respectively.
Figure 1. (a) The XRD patterns of the fly ash treated with different L/S; (b,c) SEM images of the original and washed fly ash; (dg) the X-ray mapping of the original fly ash and washed fly ash (with the main element contents shown in the inset); (h) variations in leaching concentrations of target HMs; (i) final solution pH values conducted at different temperatures (with the trace elements of the original fly ash shown in the inset of (h)). Note: L/S = 8 mL/g, washing time = 10 min, washing temperature = 25 °C, 50 °C, and 80 °C, respectively.
Sustainability 15 05873 g001
Figure 2. (ac) Variations in Pb removal efficiency and solution pH values versus bubbling time of simulated exhaust gas; (df) variations in Pb concentration versus solution pH values (the dotted lines represent the species distributions of carbonates in the aqueous solution). Note: L/S = 8 mL/g, washing time = 10 min, washing temperature = 25 °C, 50 °C, and 80 °C, respectively; gas flow rates of 10 and 15 sccm, CO2 partial pressure = 15 mol% and 33 mol%.
Figure 2. (ac) Variations in Pb removal efficiency and solution pH values versus bubbling time of simulated exhaust gas; (df) variations in Pb concentration versus solution pH values (the dotted lines represent the species distributions of carbonates in the aqueous solution). Note: L/S = 8 mL/g, washing time = 10 min, washing temperature = 25 °C, 50 °C, and 80 °C, respectively; gas flow rates of 10 and 15 sccm, CO2 partial pressure = 15 mol% and 33 mol%.
Sustainability 15 05873 g002
Figure 3. (a) ln (c/c0) versus bubbling time of simulated gas (with the kinetic constant of linear fitting pseudo-first-order equation of Pb shown in the inset); (b) the plot of ΔG0 versus T. Note: L/S = 8 mL/g, washing time = 10 min, washing temperature = 25 °C, 50 °C, and 80 °C, respectively.
Figure 3. (a) ln (c/c0) versus bubbling time of simulated gas (with the kinetic constant of linear fitting pseudo-first-order equation of Pb shown in the inset); (b) the plot of ΔG0 versus T. Note: L/S = 8 mL/g, washing time = 10 min, washing temperature = 25 °C, 50 °C, and 80 °C, respectively.
Sustainability 15 05873 g003
Figure 4. (ac) Variations in Pb species versus pH values of the washing solution in the binary system (Pb, CO32−); (df) linear sweep of the EDS spectra for Ca and Pb elements in the selected area of CaCO3 purged with different CO2 bubbling rates (10, 15 and 30 sccm); (g,h) full scan and Pb 4f of the XPS analysis for CaCO3 samples purged with different CO2 bubbling rates; (i) the removal mechanism of Pb during the CO2 bubbling process. Note L/S = 8 mL/g, washing time = 10 min, washing temperature = 25 °C; bubbling experiments: CO2 ratio as 33%, CO2 flow rates of 10, 15, 30, 45, and 61 sccm, respectively.
Figure 4. (ac) Variations in Pb species versus pH values of the washing solution in the binary system (Pb, CO32−); (df) linear sweep of the EDS spectra for Ca and Pb elements in the selected area of CaCO3 purged with different CO2 bubbling rates (10, 15 and 30 sccm); (g,h) full scan and Pb 4f of the XPS analysis for CaCO3 samples purged with different CO2 bubbling rates; (i) the removal mechanism of Pb during the CO2 bubbling process. Note L/S = 8 mL/g, washing time = 10 min, washing temperature = 25 °C; bubbling experiments: CO2 ratio as 33%, CO2 flow rates of 10, 15, 30, 45, and 61 sccm, respectively.
Sustainability 15 05873 g004
Figure 5. (a) The XRD patterns of particles obtained at different CO2 flow rates; (b) Relative contents of vaterite and calcite obtained at different CO2 flow rates; (c,d) The SEM images of aggregates composed of calcite and vaterite, respectively. Note: CO2 flow rates of 10, 15, 30, 45 and 61 sccm; L/S = 8 mL/g, washing time = 10 min, washing temperature = 25 °C.
Figure 5. (a) The XRD patterns of particles obtained at different CO2 flow rates; (b) Relative contents of vaterite and calcite obtained at different CO2 flow rates; (c,d) The SEM images of aggregates composed of calcite and vaterite, respectively. Note: CO2 flow rates of 10, 15, 30, 45 and 61 sccm; L/S = 8 mL/g, washing time = 10 min, washing temperature = 25 °C.
Sustainability 15 05873 g005
Table 1. Thermodynamic parameters during the carbonation process (at 25 °C, 50 °C, and 80 °C, respectively).
Table 1. Thermodynamic parameters during the carbonation process (at 25 °C, 50 °C, and 80 °C, respectively).
CationsTemp.Rates
(sccm)
CO2%
Ratio
KdG0
(kJ/mol)
Pb(II) 10334.622−3.795
25 °C1515127.833−12.024
3345.706−8.680
10333.185−3.113
50 °C15153.369−3.264
338.547−5.764
10332.166−2.269
80 °C15151.895−1.877
333.572−3.738
Temp.Rates
(sccm)
CO2% RatioH0
(kJ/mol)
S0
(J/mol·K)
R2
25 °C1033−10.571−230.989
50 °C1515−30.568−760.997
80 °C1533−56.742−1580.881
Table 2. Comparison of carbonation parameters reported in the available documents associated with alkaline solid waste.
Table 2. Comparison of carbonation parameters reported in the available documents associated with alkaline solid waste.
EntrySystemCaO Fraction (wt %)Reaction ParametersSolidification
Efficiency
CO2 Sequestration Efficiency **Obtained
CaCO3
Ref.
Gas–liquid–solidMSWI FA−CO2−water (China)CaO: 38.8%,Dry ash and ash with 20% H2O; CO2 content: 10%, 50%, 100%; Time: 2 hPb: 40% (final pH: 11)3% (w/w)
CO2
NA[54]
Coal FA−CO2−water (Australia)CaO: 12.5–24.8%PCO2: 3 MPa; T: 20–80 °C; L/S ratios of 0.1–0.5NA27.05 kg CO2/t FANA[55]
Circulating fluid bed (CFB) FA−CO2−steam addition (China)CaO: 28.42%,T: 300–800 °C; CO2 content: 5%, 10%, 15%, 20%, 100% (vol%); H2O content: 5%, 10%, 20%; Time: 1 hNA60 g CO2/kg FA (28.74%)NA[56]
Coal FA−Supercritical CO2 coupled with mechanical force-water (China)CaO: 25.8 wt %,T: 20–80 °C; PCO2: 1–8 Mpa; L/S = 1–300 mL/g; Time: 1–25 h; Stirring rate: 100–800 rpm; L/S = 1–300 mL/gPb: 43.4%, Cr: 98%, Cd: 60.3%42.3 mg CO2/kg FA (1 MPa); 54 mg CO2/kg FA (8 MPa)NA[57]
MSWI FA−oxy-fuel combustion flue gas−calcium carbonate oligomer regulation-water (China)CaO: 47.3% *T: 20 °C; PCO2: 0.1 Mpa; L/S = 10 L/kg; Flow rate: 200 mL/min; Time: 60 min; Stirring rate: 600 rpmPb:100%, Cu: 91.3%, Zn: 99.1%13.8%C[58]
FA washing solution−CO2 (China)CaO: 53 wt % (FA)Room temperature; Stirring rate: 63 rpm; Flow rate: 10 mL/minPb: 99%, Cu: 95%, Zn: 96%NANA[59]
Electric arc furnace steelmaking slags−CO2−NH4Cl (China)CaO: 39.04 wt %T: 12–65 °C; L/S = 20 mL/g; MW irradiation = 90–270 WNANAC+V[60]
MSWI FA−simulated exhaust gas (China)NAT: 25 -80 °C; L/S = 8 mL/g; Flow rates: 10, 15, 30, 45, and 61 sccmPb: 97.9–99.2% (final pH: 6–8)NAC+VThis study
* These data were calculated with the Ca fraction in the corresponding literature; ** Carbonation efficiency: the percentage of calcium in the FA/leachate/dust/slag converted into carbonates; FA: Fly ash; NA: not available; C: calcite CaCO3, V: vaterite CaCO3, C+V: mixture of calcite and vaterite CaCO3.
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

Wang, L.; Tang, Y.; Gong, Y.; Shao, X.; Lin, X.; Xu, W.; Zhu, Y.; Ju, Y.; Shi, L.; Kołodyńska, D. Remediation of Micro-Pollution in an Alkaline Washing Solution of Fly Ash Using Simulated Exhaust Gas: Parameters and Mechanism. Sustainability 2023, 15, 5873. https://doi.org/10.3390/su15075873

AMA Style

Wang L, Tang Y, Gong Y, Shao X, Lin X, Xu W, Zhu Y, Ju Y, Shi L, Kołodyńska D. Remediation of Micro-Pollution in an Alkaline Washing Solution of Fly Ash Using Simulated Exhaust Gas: Parameters and Mechanism. Sustainability. 2023; 15(7):5873. https://doi.org/10.3390/su15075873

Chicago/Turabian Style

Wang, Lei, Yuemei Tang, Yu Gong, Xiang Shao, Xiaochen Lin, Weili Xu, Yifan Zhu, Yongming Ju, Lili Shi, and Dorota Kołodyńska. 2023. "Remediation of Micro-Pollution in an Alkaline Washing Solution of Fly Ash Using Simulated Exhaust Gas: Parameters and Mechanism" Sustainability 15, no. 7: 5873. https://doi.org/10.3390/su15075873

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