Next Article in Journal
Ore Genesis of Shanmen Ag Deposit in Siping Area of Southern Jilin Province, NE China: Constraints from Fluid Inclusions and H-O, S, Pb Isotopes
Next Article in Special Issue
The Effect of Fly Ash Microspheres on the Pore Structure of Concrete
Previous Article in Journal
Zeta Potential of Pyrite Particles in Concentrated Solutions of Monovalent Seawater Electrolytes and Amyl Xanthate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Alumina Extraction from Coal Fly Ash via Low-Temperature Potassium Bisulfate Calcination

1
Department of Materials Science and Engineering, Liaoning Technical University, Fuxin 123000, China
2
CAS Key Laboratory of Green Process and Engineering, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China
3
Department of Environment Science and Engineering, Liaoning Technical University, Fuxin 123000, China
*
Authors to whom correspondence should be addressed.
Minerals 2019, 9(10), 585; https://doi.org/10.3390/min9100585
Submission received: 27 August 2019 / Revised: 18 September 2019 / Accepted: 25 September 2019 / Published: 27 September 2019

Abstract

:
Owing to the depletion of bauxite and increasing demand for alumina, calcination methods for extracting alumina from coal fly ash (CFA) were developed. However, these methods have disadvantages such as the need for high temperatures and the emission of toxic gases. Hence, in this study, Al2O3 was extracted from CFA via low-temperature potassium bisulfate calcination technology. Effects of the potassium bisulfate amount, calcination temperature, and calcination time on the alumina extraction efficiency were investigated using X-ray diffraction, thermal gravimetry, scanning electron microscopy, differential scanning calorimetry, and energy-dispersive spectroscopy. It was found that this technique could recover alumina efficiently, and potassium bisulfate significantly contributed to the degradation of mullite and corundum phases. Al2O3 in CFA was converted into soluble K3Al(SO4)3. With a KHSO4/Al2O3 molar ratio of 7:1, calcining temperature of 230 °C, and calcining time of 3 h, the alumina extraction efficiency reached a maximum of 92.8%. The Avrami–Erofeev equation showed the best fit with the kinetic data for the low-temperature calcination of CFA with KHSO4. The activation energy was 28.36 kJ/mol.

Graphical Abstract

1. Introduction

The discharged amount of coal fly ash (CFA) from high-temperature coal-burning power plants is expected to be 560 to 610 Mt/year based on the China Electric Power Federation forecast from 2016 to 2019 [1]; it will exceed three billion tons by 2020 in China and is behind only the tailings found in industrial solid waste [2]. Uncontrolled aggregation and discharge make CFA a main potential source of dangerous chemicals that could be released into the environment [3,4]. To enhance its utilization effectiveness, researchers made considerable efforts to study the potential of using CFA as a raw material for soil modification, production of glass/ceramics, zeolites or geopolymers, and metal extraction [5,6,7]. In particular, the CFA discharged in Shanxi and Inner Mongolia contains more alumina (40–50%) than the boundary bauxite in China and is a non-bauxite aluminum resource with high development and utilization potential [8]. The alumina industry of China imported more than 60 million tons of bauxite in 2017 [9]. Therefore, the depletion of bauxite reserves and increases in the alumina demand promoted studies on alumina extraction from high-alumina CFA in China [10].
CFA is generated at very high temperatures and consists of mullite, quartz, and some amorphous phases. Specifically, mullite is a major inactive source of Al in CFA [11]. Several processes for extracting alumina from CFA were reported and can be broadly classified into acid [12,13,14,15], alkali [16,17,18,19], and sintering processes [20]. Acid processes can effectively achieve the separation of silica and alumina in CFA [21,22]. However, acid processes inevitably cause serious equipment corrosion and, thus, require expensive acid-resistant equipment. In addition, impurities such as iron oxide and calcium oxide are difficult to remove from the leaching solution [23]. Alkali processes can achieve high alumina extraction efficiency, but the alkali solutions with high concentrations used in these processes cause a high caustic ratio in the final solution. Sintering processes are considered to have significant potential for industrial use as they are similar to the major extraction technologies currently used in the Al2O3 plants [24]. However, the use of calcination is limited by the formation of more toxic gases and the reduced activation effectiveness at high temperatures. Nevertheless, the use of additives (e.g., lime and ammonium sulfate) during heating can lower the calcination temperature and promote Al2O3 extraction [25,26]. One of the advantages of these additives is that they are widely available at low costs. There are some limitations of these processes. Alumina extraction from CFA through sintering is costly, and the residual red mud generated because of the high reaction temperature and lime addition is a new pollutant. As reported earlier, adding low amounts of lime to CFA increases the alumina dissolution to 94.5% during calcination at 1050 °C and leaching in an alkaline solution [27]. High-purity alumina was produced through sintering, leaching, NH4Al(SO4)2 precipitation, and calcination of fly ash and ammonium sulfate. With this method, the Al extraction rate under optimized conditions was 90.11% [28]. Van der Merwe et al. [29] used either ammonium sulfate or ammonium bisulfate or a mixture of the two salts as an extracting agent for recovering aluminum from ultrafine CFA during a 2-h thermal treatment process at different processing temperatures.
Although thermal activation improves alumina extraction, problems such as the need for high calcination temperatures and release of toxic gases like NH3 still exist. Therefore, to facilitate Al2O3 extraction from CFA via thermal treatment, researchers have to develop efficient methods that can decrease the calcination temperature, as well as the release of poisonous gases.
Hence, in this study, a novel thermal activation method based on potassium bisulfate (KHSO4) was established for alumina extraction from high-alumina CFA; this method was aimed at inhibiting the formation of toxic gases and reducing the calcination temperature in comparison to conventional NH4HSO4 sintering. The thermal behavior, phase conversion, and alumina extraction rates under different conditions were studied. Next, the reaction kinetics were calculated and fitted with a theoretical model. Finally, the mechanisms of CFA activation and alumina extraction were explored.

2. Experimental

2.1. Materials

CFA was collected from the Jungar Power Plant in Inner Mongolia, China, and its chemical composition is given in Table 1. Clearly, over 85 wt.% of the CFA is composed of Al2O3 and SiO2. Al2O3 accounts for up to 51.7 wt.% of CFA, which makes CFA a possible non-bauxite resource for Al2O3 generation. It also contains traces of CaO, TiO2, Fe2O3, K2O, FeO, MgO, and Na2O.
Analytical-grade KHSO4 was purchased from Nanjing Chemical Reagent Co., Ltd. (Nanjing, China). All the substances were also of analytical grade.

2.2. Procedures

The new method is schematically illustrated in Figure 1. Each time, CFA (10 g) was blended fully with KHSO4, followed by thermal treatment. The KHSO4/Al2O3 molar ratio was 4, 5, 6, 7, or 8. Then, the mixtures placed in quartz crucibles were heated at 5 °C/min in a KY-R-ST80 muffle furnace to different target temperatures ranging from 190 to 250 °C. Once the desired temperatures were reached, the corresponding crucibles were left for calcination from 1–4 h. Subsequently, the crucibles were cooled naturally. After that, the product was suspended in distilled water at a concentration of 1 g/3 mL in a beaker, which was heated to 100 °C and mixed for 30 min. The beaker was then cooled naturally. The beaker contents were filtered and rinsed with deionized water repeatedly. Then, alumina level in the leachate was detected by ethylenediamine tetraacetic acid–potassium fluoride (EDTA–KF) titrimetry [30].

2.3. Characterization

The alumina level in the filtrate obtained after washing the contents in the beaker with deionized water was detected via ethylenediamine tetraacetic acid–potassium fluoride (EDTA–KF) titrimetry [30]. The chemical compositions of the CFA, calcined products, and residue were measured via X-ray fluorescence analysis (XRF, Axios, PANalytical, Almelo, The Netherlands).
Crystalline phases of the CFA and calcined products were investigated via X-ray diffraction (XRD, 6100, Shimadzu, Kyoto, Japan) using Cu Kα radiation with the following parameters: acceleration voltage = 40 kV, current = 30 mA, scanning angles ranging from 5° to 60°, and scanning speed of 2°/min.
Thermal gravimetry–differential scanning calorimetry (TG–DSC) was conducted using an STA449C integrated instrument (Netzsch, Selb, Germany) under N2 flow; samples weighing 10 mg were heated from 25 to 700 °C at a heating rate of 5 °C/min.
Morphologies of the CFA, calcined products, and residue were characterized via scanning electron microscopy (SEM, 7500F, JEOL, Tokyo, Japan), and chemical compositions were analyzed via energy-dispersive spectroscopy (EDS).
The Al concentration in the filtrates was detected using EDTA–KF titrimetry with alumina as the standard. The alumina extraction rate (ω(Al2O3)) was computed as follows:
ω ( Al 2 O 3 ) = m F ( Al 2 O 3 ) m F A ( Al 2 O 3 ) ,
where mFA(Al2O3) is the Al2O3 mass in the fly ash, and mF(Al2O3) is that in the filtrate. Each experiment was carried out thrice, and the mean was calculated. The uncertainty for alumina was set at ±2%.

3. Results and Discussion

3.1. TG–DSC Results

Figure 2 presents the weight loss and heat flow data of pure KHSO4 and that of the fly ash with KHSO4. The DSC curve of pure KHSO4 showed several exothermic and endothermic peaks in the range of 100–750 °C (Figure 2a). The peaks at 199 °C and 217 °C correspond to the melting and phase conversion of KHSO4 to K2S2O7 with a mass loss of approximately 6.35%. From 350 °C to 550 °C, KHSO4 or K2S2O7 slowly transformed to K2SO4 with a mass loss of approximately 6.98%. The rapid degradation when the temperature reached approximately 590 °C indicates the completion of the phase transition.
The DSC curve of the fly ash with KHSO4 showed endothermic peaks from 71 to 226 °C (Figure 2b). The first peak at 71 °C indicates a mass loss of 0.4% owing to the elimination of physically adsorbed water. The second peak at 180–270 °C indicates a mass loss of 3.15% and coexists with two peaks at 206 and 226 °C, which correspond to the melting of KHSO4 and transformation of minor KHSO4 into K2S2O7 and active reaction of KHSO4 with CFA, respectively [31]; the reaction is shown in Equation (2). The third mass loss of 6.37% at 400–570 °C, together with another peak, corresponds to the burning of unburnt carbon in CFA [32,33]. The exothermic and endothermic peaks from 350 °C to 550 °C are not present in Figure 2b, suggesting that KHSO4 completely reacted with CFA.
2 KHSO 4 ( s ) K 2 S 2 O 7 ( s ) + H 2 O .

3.2. Effects of Calcination Parameters on Alumina Extraction Efficiency

Figure 3, Figure 4 and Figure 5 show the effects of the calcination temperature, KHSO4/Al2O3 molar ratio, and calcination time on the efficiency of alumina extraction from CFA, respectively. The error bar represents the standard error of the mean when comparing all values. As the temperature increased, the extraction efficiency improved gradually, reached a maximum of 92.8% at 230 °C, and then increased gently. At KHSO4/Al2O3 molar ratios < 6, increasing KHSO4 dosage most significantly improved the alumina extraction efficiency, which reached 88.3% at a molar ratio of six. The alumina extraction efficiency increased up to 92.8% at a ratio of seven, but slightly reduced to 92.7% when the molar ratio further increased to eight. Moreover, the alumina extraction efficiency nearly reached its maximum value after 3 h of reaction (Figure 5) but did not significantly improve with further prolonging of the reaction time.

3.3. Decomposition Mechanism

To explore how KHSO4 addition activated CFA and the underlying mechanism, we investigated the crystallographic structures of the CFA, calcined products, and residue via XRD.
As shown in Figure 6, the main crystal phase in the calcined products was K3Al(SO4)3, without mullite or corundum. This result suggests that KHSO4 actively reacted with the CFA, which considerably accelerated the degradation of mullite and corundum. The low-temperature thermal decomposition of CFA by KHSO4 can be understood through the following reactions:
18 KHSO 4 ( l ) + 3 Al 2 O 3 2 SiO 2 ( m u l l i t e , s ) 6 K 3 Al ( SO 4 ) 3 ( s ) + 2 SiO 2 ( s ) + 9 H 2 O ,
6 KHSO 4 ( l ) + Al 2 O 3 ( c o r u n d u n , s ) 2 K 3 Al ( SO 4 ) 3 ( s ) + 3 H 2 O ,
9 K 2 S 2 O 7 ( s ) + 3 Al 2 O 3 2 SiO 2 ( m u l l i t e , s ) 6 K 3 Al ( SO 4 ) 3 ( s ) + 2 SiO 2 ( s ) ,
3 K 2 S 2 O 7 ( s ) + Al 2 O 3 ( c o r u n d u m , s ) 2 K 3 Al ( SO 4 ) 3 ( s ) .
As shown in Figure 7, the dominant crystal phases of CFA were mullite and corundum, which correspond to peaks at 2θ = 16.5°, 25.9°, 26.2°, 35.2°, and 40.8° and at 25.6°, 37.8°, 43.4°, and 64.5°, respectively. The evident bulge at 2θ = 22° corresponds to tridymite and cristobalite [34]. After calcination at 190 °C, the mullite (16.5°, 25.9°, 26.2°, 35.2°, and 40.8°) and corundum peaks (25.6°, 37.8°, 43.4°, and 64.5°) gradually weakened in intensity, while the slight peaks of amorphous silicon oxide from 18° to 24° became stronger. This was because the mullite and corundum in CFA containing KHSO4 degraded under these conditions. These results are consistent with the alumina extraction rate of 54.7% given in Figure 3. Furthermore, at 210 °C, the corundum (25.6°, 37.8°, 43.4°, and 64.5°) peaks disappeared, indicating the nearly complete decomposition of corundum. Nevertheless, some mullite peaks remained. Accordingly, the Al2O3 extraction rate at 340 °C improved to 80.6%. At 230 °C, amorphous quartz also showed peaks from 18° to 24°, suggesting that mullite also fully degraded, and the alumina extraction efficiency increased to 92.8%. At 250 °C, 93.1% of alumina was extracted from CFA, suggesting that the diffraction peaks of the residue changed little compared to those at 230 °C.
At a KHSO4/Al2O3 molar ratio of four, the typical peaks of mullite and corundum weakened in intensity (Figure 8). When the ratio increased to five, these characteristic peaks weakened further. Moreover, the peaks of mullite at 2θ = 26.2° and those of corundum at 25.6° and 37.8° were hardly visible. Additionally, the alumina extraction rates were below 70%, indicating the incomplete reactions of KHSO4 with mullite and corundum. At a molar ratio of six, the main peaks of mullite at 2θ = 16.5° and 25.9° and those of corundum disappeared; however, the peaks of mullite at 26.2°, 35.2°, and 40.8° reduced in intensity, and the alumina extraction efficiency nearly reached a maximum of 88.3%. At KHSO4/Al2O3 molar ratios > 7, neither the mullite nor the corundum peaks could be found, and the alumina extraction efficiency slightly increased, implying that the NaHSO4 dosage could be excessive. Therefore, 7:1 was considered the optimal KHSO4/Al2O3 molar ratio.
As shown in Figure 9, after 1 h of calcination, the peaks of mullite (16.5°, 25.9°, 26.2°, 35.2°, and 40.8°) and corundum (43.4° and 64.5°) gradually reduced in intensity, while the peaks of corundum at 25.6° and 37.8° disappeared. After 2 h of calcination, the characteristic peaks of mullite (26.2°, 35.2°, and 40.8°) and corundum (43.4° and 64.5°) weakened in intensity, while the other mullite peaks (16.5° and 25.9°) disappeared. As the calcination time increased to 3 h, the characteristic peaks of neither mullite nor corundum could be found in the spectra. Considering the results presented in Figure 5 and Figure 9, it was concluded that the mineral phases of mullite and corundum in CFA were almost fully degraded by KHSO4 during low-temperature calcination. After 4 h of calcination, the characteristic peaks did not significantly change in terms of their intensity, suggesting the reaction between CFA and KHSO4 was not completed until 3 h. Thus, 3 h was considered the optimal calcination time for ensuring the complete reaction.

3.4. Analysis of Reaction Kinetics Using Kinetic Model

Chemical reactions involving solid reactants occur at the reactant interface to form compounds [35]. The reactions do not continue without the diffusion of products and will proceed at the interface of one or both reactants. As the mechanisms of these reactions depend on the test conditions and on the properties of the solids, the overall process rate may be controlled by chemical kinetics or transport processes.
The kinetic models considered in recent studies are listed in Table 2 [36]. Considering the reaction data obtained in this study, models (1)–(5) were considered highly suitable for describing these reactions, with a correlation coefficient higher than 0.98. This implies that the reaction is limited by the diffusion of reactants through the product layer. It is also obvious that the kinetic data show the best fit with the Avrami–Erofeev equation at all temperatures.
The Avrami–Erofeev equation based on the concept of nucleation and nuclei growth [37] can successfully explain the leaching kinetics of various solid–liquid reactions [38,39]. In this study, the kinetics of the reaction between CFA and KHSO4 were analyzed using the Avrami–Erofeev equation, given as follows:
ln ( 1 α ) = ( k t ) n ,
where α, k, and t are the alumina extraction efficiency, rate constant, and time of the reaction, respectively, and n is the mechanism-specific Avrami exponent. Equation (7) can be linearly converted to Equation (8).
ln [ ln ( 1 α ) ] = n ln t + n ln k .
The reaction kinetics were analyzed using the data acquired at 210, 220, and 230 °C, and the results are presented in Table 3. The exponent n is equal to the slope of the straight line in the plot of ln[−ln(1 − α)] versus ln(t), and the rate constant k can be calculated from the intercepts, as shown in Figure 10.
The apparent activation energy was calculated using the Arrhenius equation based on the results presented in Table 2. The values of R2 suggest that this equation reasonably fits with the experimental data. The apparent activation energy, equal to the slope of the straight line in the plot of the natural logarithm of the reaction rate versus 1000/T (Figure 11), was 28.36 kJ/mol with R2 equal to 0.98666. These kinetic calculation results will be highly valuable for performing parameter adjustments and scaling up alumina production.

3.5. SEM and EDS Results

The SEM and EDS images of CFA and the residue are shown in Figure 12 and Figure 13, respectively. The raw CFA comprised size-variable spherical bodies, having a smooth surface, composed of mullite or aluminosilicate glass resulting from the reactions between alumina and SiO2 during coal combustion (Figure 12). These results agree with those reported in earlier studies [40,41]. EDS analysis showed that the major elements in the CFA were Si, Al, Ca, and O constituting different compounds in CFA. Al was mainly related with Si. The particle diameter of the residue was approximately 2–5 μm (Figure 13). The general structure of CFA was damaged, and the mullite on the spherical surface was fully degraded during low-temperature calcination. EDS showed that the proportions of major elements in these particles differed significantly from those presented in Figure 12. EDS mapping showed the major component was amorphous silica, which is in accordance with the XRD images. Table 4 presents the chemical composition of the residue after alumina extraction. Compared to the chemical composition of CFA in given Table 1, the Al2O3 content was reduced from 51.70% to 13.88%, and SiO2 content increased from 36.30% to 70.32%. This shows that most of the aluminum components in CFA were extracted, while the silica components were retained in the residue. This eliminates the separation process of silica and aluminum that is required in alkali processes and reduces slag production compared to the alkali or alkali lime calcination processes; therefore, the discharge of solid waste and the resulting environmental pollution are reduced. The SiO2 content in the residue is high, and its activity is high as well. It can be used as a raw material to prepare silicon-based products, such as zeolite, tobermorite, xonotlite, and geopolymers [5,42,43,44].
Figure 14 shows the SEM images of CFA and the residue after calcination at 230 °C at different durations. After 1 h of calcination, the structure of CFA was eroded by KHSO4, showing holes and depressions to different degrees, as well as rough spherical surfaces (Figure 14); thus, the activation of KHSO4 started to destroy the surface structure of CFA. When the reaction proceeded for 2 h, the spherical particles were eroded further, and they started to break. The entrapped globules were exposed and released, and they continued to react with KHSO4. At a calcination time of 3 h, the spherical particles were completely destroyed, and the entrapped globules were all released. The Si–O–Al bonds in mullite were broken so that alumina could be extracted.

4. Conclusions

(1) A new method for Al2O3 extraction from CFA was proposed by adding KHSO4 at low calcination temperatures. This method offers the following advantages: low energy consumption, no emission of toxic gases, high extraction efficiency, economical, and environmentally friendly.
(2) The reaction between CFA and KHSO4 formed soluble K3Al(SO4)3. The calcination temperature, KHSO4/Al2O3 molar ratio, and duration of calcination considerably impacted the Al2O3 extraction rate. The extraction efficiency of alumina from CFA reached up to 92.8% with a KHSO4/Al2O3 molar ratio of 7:1 when calcination was carried out at 230 °C for 3 h.
(3) The Avrami–Erofeev equation showed the best fit with the kinetic data for low-temperature calcination of CFA with KHSO4. The activation energy was 28.36 kJ/mol.
(4) XRD results showed that the mullite and corundum in CFA almost decomposed under the abovementioned optimum conditions, and the residue was mainly composed of SiO2. SEM results showed that the spherical CFA particles decomposed completely during low-temperature calcination, and EDS mapping also showed that the residue was primarily composed of SiO2.

Author Contributions

Conceptualization, C.G.; Methodology, C.G. and J.Z.; Validation, J.Z. and S.M.; Investigation J.Y. and K.W.; Writing-Original Draft Preparation, C.G. and J.Z.; Writing-Review & Editing, J.Z.; Funding Acquisition, C.G., J.Z., S.M. and J.Y.

Funding

The authors are grateful for the Liaoning Provincial College Basic Research Program of Liaoning Education Department (No. LJ2017QL004), the industrial public relations and industrialization guidance plan of Liaoning Provincial Department of Science and Technology (2019JH8/10100092), the Undergraduate Training Program for Innovation and Entrepreneurship of Liaoning Provincial (No. 201810147047 and 201810147088), the National Natural Science Foundation of China (No. 51404136), the Major Science and Technology Program of the Inner Mongolia Autonomous Region (Preparation and application demonstration of fly ash based soil conditioner), and the Natural Science Foundation Guide Project of Liaoning Province (20180551120).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, S.Y.; Dai, S.F.; Finkelman, R.B.; Graham, I.T.; French, D.; Hower, J.C.; Li, X.W. Leaching characteristics of alkaline coal combustion by-products: A case study from a coal-fired power plant, Hebei province, China. Fuel 2019, 255, 115710. [Google Scholar] [CrossRef]
  2. Yao, Z.T.; Xia, M.S.; Sarker, P.K.; Chen, T. A review of the alumina recovery from coal fly ash with a focus in China. Fuel 2014, 120, 74–85. [Google Scholar] [CrossRef]
  3. Blissett, R.S.; Rowson, N.A. A review of the multi-component utilisation of coal fly ash. Fuel 2012, 97, 1–23. [Google Scholar] [CrossRef]
  4. Ahmaruzzaman, M. A review on the utilization of fly ash. Prog. Energy Combust. Sci. 2010, 36, 327–363. [Google Scholar] [CrossRef]
  5. Zou, J.J.; Guo, C.B.; Zhou, X.Q.; Sun, Y.J.; Yang, Z. Sorption capacity and mechanism of Cr3+ on tobermorite derived from fly ash acid residue and carbide slag. Colloid Surf. A Physicochim. Eng. Asp. 2018, 538, 825–833. [Google Scholar] [CrossRef]
  6. Nath, P.; Sarker, P.K. Flexural strength and elastic modulus of ambient-cured blended low-calcium fly ash geopolymer concrete. Constr. Build. Mater. 2017, 130, 22–31. [Google Scholar] [CrossRef]
  7. Bukhari, S.S.; Behin, J.; Kazemian, H.; Rohani, S. Conversion of coal fly ash to zeolite utilizing microwave and ultrasound energies: A review. Fuel 2015, 140, 250–266. [Google Scholar] [CrossRef]
  8. Guo, C.B.; Zou, J.J.; Wei, C.D.; Jiang, Y.S. Comparative study on extracting alumina from circulating fluidized-bed and pulverized-coal fly ashes through salt activation. Energy Fuel 2013, 27, 7868–7875. [Google Scholar] [CrossRef]
  9. Wang, X.W. China’s Bauxite Resource Product Demand Forecasting and Countermeasure Study; China University of Geosciences: Beijing, China, 2018. [Google Scholar]
  10. Guo, Y.; Li, Y.; Cheng, F.; Wang, M.; Wang, X. Role of additives in improved thermal activation of coal fly ash for alumina extraction. Fuel Process. Technol. 2013, 110, 114–121. [Google Scholar] [CrossRef]
  11. Gomes, S.; Francois, M. Characterization of mullite in silicoaluminous fly ash by XRD, TEM, and 29Si MAS NMR. Cem. Concr. Res. 2000, 30, 175–181. [Google Scholar] [CrossRef]
  12. Bai, G.H.; Qiao, Y.H.; Shen, B.; Chen, S.L. Thermal decomposition of coal fly ash by concentrated sulfuric acid and alumina extraction process based on it. Fuel Process. Technol. 2011, 92, 1213–1219. [Google Scholar] [CrossRef]
  13. Shemi, A.; Ndlovu, S.; Sibanda, V.; van Dyk, D.V. Extraction of alumina from coal fly ash using an acid leach-sinter-acid leach technique. Hydrometallurgy 2015, 157, 348–355. [Google Scholar] [CrossRef]
  14. Nayak, N.; Panda, C.R. Aluminium extraction and leaching characteristics of Talcher Thermal Power Station fly ash with sulphuric acid. Fuel 2010, 89, 53–58. [Google Scholar] [CrossRef]
  15. Ji, H.M.; Lu, H.X.; Hao, X.G.; Wu, P. High purity alumina powders extracted from fly ash by the calcining-leaching process. J. Chin. Ceram. Soc. 2007, 35, 1657–1660. [Google Scholar]
  16. Sun, Z.H.; Li, H.Q.; Bao, W.J.; Wang, C.Y. Mineral phase transition of desilicated high alumina fly ash with alumina extraction in mixed alkali solution. Int. J. Miner. Process. 2016, 153, 109–117. [Google Scholar] [CrossRef]
  17. Yang, Z.; Sun, J.; Zhang, Z.; Ye, J.; Miao, R. Using silicate-calcium slag generated in process of extracting alumina from fly ash as cement admixture. Chin. J. Environ. Eng. 2014, 8, 3989–3995. [Google Scholar]
  18. Ma, Y.; Hu, J.; Ye, G. The effect of activating solution on the mechanical strength, reaction rate, mineralogy, and microstructure of alkali-activated fly ash. J. Mater. Sci. 2012, 47, 4568–4578. [Google Scholar] [CrossRef] [Green Version]
  19. Louet, N.; Gonon, M.; Fantozzi, G. Influence of the amount of Na2O and SiO2 on the sintering behavior and on the microstructural evolution of a Bayer alumina powder. Ceram. Int. 2005, 31, 981–987. [Google Scholar] [CrossRef]
  20. Guo, C.B.; Zou, J.J.; Jiang, Y.S.; Huang, T.P.; Cheng, Y.; Wei, C.D. Thermal activation of coal fly ash by sodium hydrogen sulfate for alumina extraction. J. Mater. Sci. 2014, 49, 4315–4322. [Google Scholar] [CrossRef]
  21. Valeev, D.; Kunilova, I.; Alpatov, A.; Mikhailova, A.; Goldberg, M.; Kondratiev, A. Complex utilisation of ekibastuz brown coal fly ash: Iron & carbon separation and aluminum extraction. J. Clean. Prod. 2019, 218, 192–201. [Google Scholar]
  22. Ma, Z.B.; Zhang, S.; Zhang, H.R.; Cheng, F.Q. Novel extraction of valuable metals from circulating fluidized bedderived high-alumina fly ash by acidealkaliebased alternate method. J. Clean. Prod. 2019, 230, 302–313. [Google Scholar] [CrossRef]
  23. Matjie, R.H.; Bunt, J.R.; Van Heerden, J.H.P. Extraction of alumina from coal fly ash generated from a selected low rank bituminous South African coal. Miner. Eng. 2005, 18, 299–310. [Google Scholar] [CrossRef]
  24. Liu, K.; Xue, J.L.; Zhu, J. Extracting alumina from coal fly ash using acid sintering-leaching process. Light Met. 2012, 201–206. [Google Scholar]
  25. Bai, G.H.; Wang, T.S.; Li, P.C.; Lv, L.; He, C.L. Separation of Gallium from aluminum in fly ash digested solution by soda-lime sintering process. Adv. Mater. Res. 2012, 560, 474–481. [Google Scholar] [CrossRef]
  26. Doucet, F.J.; Mohamed, S.; Neyt, N.; Castleman, B.A.; van der Merwe, E.M. Thermochemical processing of a South African ultrafine coal fly ash using ammonium sulphate as extracting agent for aluminium extraction. Hydrometallurgy 2016, 166, 174–184. [Google Scholar] [CrossRef] [Green Version]
  27. Jiang, Z.Q.; Ma, H.W.; Yang, J.; Ma, X.; Yuan, J.Y. Thermal decomposition mechanism of desilication coal fly ash by low-lime sinter method for alumina extraction. Ferroelectrics 2015, 486, 143–155. [Google Scholar] [CrossRef]
  28. Park, H.C.; Park, Y.J.; Stevens, R. Synthesis of alumina from high purity alum derived from coal fly ash. Mater. Sci. Eng. A 2004, 367, 166170. [Google Scholar] [CrossRef]
  29. Van der Merwea, E.M.; Graya, C.L.; Castlemana, B.A.; Mohamed, S.; Krugerc, R.A.; Doucet, F.J. Ammonium sulphate and/or ammonium bisulphate as extracting agents for the recovery of aluminium from ultrafine coal fly ash. Hydrometallurgy 2017, 171, 185–190. [Google Scholar] [CrossRef] [Green Version]
  30. Ministry of Public Health. Pharmacopoeia of PRC; Chemical Industry Press: Beijing, China, 1995. [Google Scholar]
  31. Eriksen, K.M.; Fehrmann, R.; Hatem, G.; Gaune-Escard, M.; Lapina, O.B.; Mastikhin, V.M. Conduvtivity, NMR, thermal measurements and phase diagram of the K2S2O7-KHSO4 System. J. Phys. Chem. 1996, 100, 10771–10778. [Google Scholar] [CrossRef]
  32. Sánchez, R.M.T.; Basaldella, E.I.; Marco, J.F. The effect of thermal and mechanical treatments: Characterization by XPS and IEP measurements. J. Colloid Interface Sci. 1999, 215, 339–344. [Google Scholar] [CrossRef]
  33. Payá, J.; Monzó, J.; Borrachero, M.V.; Perris, E.; Amahjour, F. Thermogravimetric methods for determining carbon content in fly ashes. Cem. Concr. Res. 1998, 28, 675–686. [Google Scholar] [CrossRef]
  34. Zhang, Z.J.; Sun, J.M.; Yao, Q. Research on the extraction amorphous SiO2 from high-aluminium fly ash. Acta Miner. Sin. 2007, 27, 137–142. [Google Scholar]
  35. Padilla, R.; Sohn, H.Y. Sintering Kinetics and Alumina Yield in lime-soda sinter process for alumina from coal wastes. Metall. Trans. B 1985, 16, 385–395. [Google Scholar] [CrossRef]
  36. Rogozhnikov, D.A.; Shoppert, A.A.; Dizer, O.A.; Karimov, K.A.; Rusalev, R.E. Leaching Kinetics of Sulfides from Refractory Gold Concentrates by Nitric Acid. Metals 2019, 9, 465. [Google Scholar] [CrossRef]
  37. Avrami, M. Kinetics of phase change. II transformation-time relations for random distribution of nuclei. J. Chem. Phys. 1940, 8, 212–224. [Google Scholar]
  38. Rodriguez, M.; Quiroga, O.; Ruiz, M.D.C. Kinetic study of ferrocolumbite dissolution in hydrofluoric acid medium. Hydrometallurgy 2007, 85, 87–94. [Google Scholar] [CrossRef]
  39. Kahruman, C.; Yusufoglu, I. Leaching kinetics of synthetic CaWO4 in HCl solutions containing H3PO4 as chelating agent. Hydrometallurgy 2006, 81, 182–189. [Google Scholar] [CrossRef]
  40. Peng, M.; Ruan, X.Y. Study on both shape and chemical composition at the surface of fly ash by scanning electron microscope, focused ion beam, and field emission-scanning electron microscope. Chin. J. Anal. Chem. 2004, 32, 1196–1198. [Google Scholar]
  41. Kutchko, B.G.; Kim, A.G. Fly ash characterization by SEM–EDS. Fuel 2006, 85, 2537–2544. [Google Scholar] [CrossRef]
  42. Zou, J.J.; Guo, C.B.; Wei, C.D.; Jiang, Y.S. Dynamic hydrothermal synthesis of xonotlite from acidextracting residues of circulating fluidized bed fly ash. Res. Chem. Intermed. 2016, 2, 519–530. [Google Scholar] [CrossRef]
  43. Zou, J.J.; Guo, C.B.; Jiang, Y.S.; Wei, C.D.; Li, F.F. Structure, morphology and mechanism research on synthesizing xonotlite fiber from acid-extracting residues of coal fly ash and carbide slag. Mater. Chem. Phys. 2016, 172, 121–128. [Google Scholar] [CrossRef]
  44. Kaze, C.R.; Djobo, J.N.Y.; Nana, A.; Tchakoute, H.K.; Kamseu, E.; Melo, U.C.; Leonelli, C.; Rahier, H. Effect of silicate modulus on the setting, mechanical strength and microstructure of iron-rich aluminosilicate (laterite) based-geopolymer cured at room temperature. Ceram. Int. 2018, 44, 21442–21450. [Google Scholar] [CrossRef]
Figure 1. Flow diagram of the new calcination method for alumina extraction from coal fly ash (CFA).
Figure 1. Flow diagram of the new calcination method for alumina extraction from coal fly ash (CFA).
Minerals 09 00585 g001
Figure 2. Thermal gravimetry–differential scanning calorimetry (TG–DSC) curves: (a) pure KHSO4 and (b) CFA blended with KHSO4.
Figure 2. Thermal gravimetry–differential scanning calorimetry (TG–DSC) curves: (a) pure KHSO4 and (b) CFA blended with KHSO4.
Minerals 09 00585 g002
Figure 3. Effects of calcination temperature on alumina extraction from CFA at a reaction time of 3 h and KHSO4/Al2O3 molar ratio of seven.
Figure 3. Effects of calcination temperature on alumina extraction from CFA at a reaction time of 3 h and KHSO4/Al2O3 molar ratio of seven.
Minerals 09 00585 g003
Figure 4. Effects of KHSO4/Al2O3 molar ratio on alumina extraction from CFA at a calcination temperature of 230 °C and time of 3 h.
Figure 4. Effects of KHSO4/Al2O3 molar ratio on alumina extraction from CFA at a calcination temperature of 230 °C and time of 3 h.
Minerals 09 00585 g004
Figure 5. Effects of calcination time on alumina extraction from CFA at 230 °C and a KHSO4/Al2O3 molar ratio of seven.
Figure 5. Effects of calcination time on alumina extraction from CFA at 230 °C and a KHSO4/Al2O3 molar ratio of seven.
Minerals 09 00585 g005
Figure 6. X-ray diffraction (XRD) patterns of products calcined at 230 °C with a KHSO4/Al2O3 molar ratio of seven for 3 h.
Figure 6. X-ray diffraction (XRD) patterns of products calcined at 230 °C with a KHSO4/Al2O3 molar ratio of seven for 3 h.
Minerals 09 00585 g006
Figure 7. XRD spectra of CFA and residue changing with calcination temperatures.
Figure 7. XRD spectra of CFA and residue changing with calcination temperatures.
Minerals 09 00585 g007
Figure 8. XRD spectra of CFA and residue at different KHSO4/Al2O3 molar ratios.
Figure 8. XRD spectra of CFA and residue at different KHSO4/Al2O3 molar ratios.
Minerals 09 00585 g008
Figure 9. XRD spectra of CFA and residue at different time periods of calcination.
Figure 9. XRD spectra of CFA and residue at different time periods of calcination.
Minerals 09 00585 g009
Figure 10. Kinetic plots obtained using the Avrami–Erofeev equation at 210, 220, and 230 °C.
Figure 10. Kinetic plots obtained using the Avrami–Erofeev equation at 210, 220, and 230 °C.
Minerals 09 00585 g010
Figure 11. Plots of −ln(k) * R versus 1000/T to determine activation energy.
Figure 11. Plots of −ln(k) * R versus 1000/T to determine activation energy.
Minerals 09 00585 g011
Figure 12. SEM and energy-dispersive spectroscopy (EDS) images of CFA.
Figure 12. SEM and energy-dispersive spectroscopy (EDS) images of CFA.
Minerals 09 00585 g012
Figure 13. SEM and EDS images of residue.
Figure 13. SEM and EDS images of residue.
Minerals 09 00585 g013
Figure 14. SEM images of CFA and residue after calcination at 230 °C for different durations: (a) CFA; (b) 1 h; (c) 2 h; (d) 3 h.
Figure 14. SEM images of CFA and residue after calcination at 230 °C for different durations: (a) CFA; (b) 1 h; (c) 2 h; (d) 3 h.
Minerals 09 00585 g014
Table 1. Composition of the coal fly ash (CFA) (wt.%).
Table 1. Composition of the coal fly ash (CFA) (wt.%).
ComponentAl2O3SiO2CaOTiO2Fe2O3K2OFeOMgONa2OLOI a
Content51.7036.302.681.751.280.340.300.240.092.21
a Loss on ignition.
Table 2. Fitting results obtained using several kinetic models.
Table 2. Fitting results obtained using several kinetic models.
Limiting StepEquationR2
190 °C210 °C230 °C
1. Deformation of Avrami–Erofeev equation ln [ ln ( 1 α ) ] = n ln t + n ln k 0.995730.996440.99755
2. Diffusion through the product layer (sp) 1 3 ( 1 α ) 2 / 3 + 2 ( 1 α ) = k t 0.988590.983590.98212
3. Diffusion through the product layer (pp) α 2 = k t 0.985000.988590.98627
4. Diffusion through the product layer (cp) α + ( 1 α ) ln ( 1 α ) = k t 0.990100.989090.98975
5. Diffusion through the liquid film (sp) α 2 = k t 0.897870.903750.89545
6. Surface chemical reactions (cp) α + ( 1 α ) 1 / 2 = k t 0.953300.965540.97435
7. Surface chemical reactions (sp) α + ( 1 α ) 1 / 3 = k t 0.967940.979940.99006
8. New shrinking core model 1 / 3 ln ( 1 α ) + [ ( 1 α ) 1 / 3 1 ] = k t 0.943350.895640.84721
sp—spherical particles, pp—plate particles, cp—cylindrical particles, k—a chemical constant, α—the degree of alumina extraction efficiency, and t—the leaching time.
Table 3. Reaction kinetics calculated using data acquired at 210, 220, and 230 °C.
Table 3. Reaction kinetics calculated using data acquired at 210, 220, and 230 °C.
Temperature (°C)Order, nRate Constant, kR2, Linear Regression Fit
2100.87160.0101066490.99573
2200.948720.0112158290.99644
2301.083140.0133996520.99755
Table 4. Composition of the residue (wt.%).
Table 4. Composition of the residue (wt.%).
ComponentAl2O3SiO2CaOTiO2Fe2O3K2OFeOMgONa2OLOI a
Content13.8870.321.932.121.151.180.180.220.197.32
a Loss on ignition.

Share and Cite

MDPI and ACS Style

Guo, C.; Zou, J.; Ma, S.; Yang, J.; Wang, K. Alumina Extraction from Coal Fly Ash via Low-Temperature Potassium Bisulfate Calcination. Minerals 2019, 9, 585. https://doi.org/10.3390/min9100585

AMA Style

Guo C, Zou J, Ma S, Yang J, Wang K. Alumina Extraction from Coal Fly Ash via Low-Temperature Potassium Bisulfate Calcination. Minerals. 2019; 9(10):585. https://doi.org/10.3390/min9100585

Chicago/Turabian Style

Guo, Chunbin, Jingjing Zou, Shuhua Ma, Jianlin Yang, and Kehan Wang. 2019. "Alumina Extraction from Coal Fly Ash via Low-Temperature Potassium Bisulfate Calcination" Minerals 9, no. 10: 585. https://doi.org/10.3390/min9100585

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