Next Article in Journal
Coal Sludge Permeability Assessment Based on Rowe Cell Consolidation, and Filtration Investigations
Previous Article in Journal
Process Mineralogy of the Tailings from Llallagua: Towards a Sustainable Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Extraction of Aluminum and Iron Ions from Coal Gangue by Acid Leaching and Kinetic Analyses

1
School of Chemical Engineering and Technology, China University of Mining and Technology, Xuzhou 221016, China
2
Guizhou Provincial Key Laboratory of Coal Clean Utilization, School of Chemistry and Materials Engineering, Liupanshui Normal University, Liupanshui 553004, China
*
Author to whom correspondence should be addressed.
Minerals 2022, 12(2), 215; https://doi.org/10.3390/min12020215
Submission received: 6 January 2022 / Revised: 26 January 2022 / Accepted: 1 February 2022 / Published: 7 February 2022

Abstract

:
Extracting valuable elements from coal gangue is an important method for the utilization of coal gangue. In order to obtain the suitable technological conditions and the acid leaching kinetic model of leaching aluminum and iron ions from high-iron and low-aluminum coal gangue, the effects of calcination temperature, calcination time, and acid types on the leaching results of aluminum and iron ions are studied. The results show that when the gangue is calcined at 675 °C for 1 h, then the calcined gangue powder is leached by 6 mol/L hydrochloric acid at 93 °C for 4 h, the leaching ratio of iron ions is more than 90%, and that of aluminum ions is more than 60%. Furthermore, the acid leaching kinetic equations at 30 °C, 50 °C, 70 °C, and 90 °C are studied by three kinetic models, and the apparent activation energies of the reactions are calculated by the Arrhenius formula. The results show that the leaching behavior of aluminum and iron ions conformed to the “mixing control” model equation: “(1 − x)−1/3 − 1 + 1/3ln(1 − x) = kt + b”. The apparent activation energies of aluminum and iron ions are 55.5 kJ/mol and 55.8 kJ/mol, respectively. All these indicate that the acid leaching process is controlled by the “mixing control”.

1. Introduction

Coal gangue is the waste produced by the coal industry [1], which accounts for 10–20% of the raw coal output [2]. Due to the massive accumulation of coal gangue, it has polluted the atmosphere, water, and soil [3], occupied land, and affected the safety of the ecological environment [4]. The resource utilization of coal gangue is conducive to decreasing its ecological damage and enhancing industrial added value. The common ways to utilize gangue include mine backfilling [5], brick making [6,7], road paving [8], cement making [9], etc. However, these methods cannot make use of coal gangue for high-value-added applications, so it is imperative to research a method of efficient development and comprehensive utilization of coal gangue [10,11].
High-value-added applications of coal gangues include the preparation of molecular sieves [12], the preparation of adsorbents [13], the extraction of elements such as aluminum and silicon [14], and the preparation of products using extracted materials such as Al(OH)3 [15], Al2O3 [16,17], AlCl3 [18], A12(SO4)3 [19], polyaluminium chloride [20], Na2SiO3, and SiO2 [21]. Most coal gangues used in these studies have low iron content, so the purity and whiteness of the prepared products are high.
The increase in iron content in coal gangue will increase the probability of iron entering the products, which will affect its application fields, and even make it difficult to carry out fine processing directly [22]. In Liupanshui, Guizhou Province, China, more than 10 million tons of coal gangue are produced every year, and coupled with years of accumulation, the current amount of coal gangue is more than 180 million tons. The components of coal gangue in this area are different from those in other areas, and its main feature is that it is rich in iron. Therefore, it cannot be used in areas requiring product purity and whiteness.
Generally, aluminum and iron elements from coal gangue can be extracted to prepare Al(OH)3, Al2O3, Fe2O3, and composite coagulant [23,24]. There are various methods for the extraction of aluminum and iron ions, such as acid leaching and alkali leaching [25,26,27,28].
With these methods, the extraction ratios of aluminum and iron elements are affected not only by the process conditions, but also by the kinds of minerals. However, the components of gangue vary from place to place, so these existing technologies cannot be copied and applied directly. Therefore, it is necessary to explore the optimal leaching process conditions.
In this work, the components, phases, morphology, and the thermal–chemical properties of the coal gangue from Liupanshui are analyzed by XRF, XRD, SEM, and TG-DSC, respectively. The suitable process conditions of extracting aluminum and iron ions are obtained. Three control models in the “unreacted core shrinkage model” (USCM) are used to fit the leaching kinetics of aluminum and iron ions in the acid leaching process of hydrochloric acid. The results show that “the mixed control” model can explain the leaching kinetics of coal gangue samples effectively.

2. Experimental Section

2.1. Materials

The coal gangue came from Wangjiazhai Coal Mine of Liupanshui, Guizhou Province, China, which was made into powder of less than 160 mesh. Sulfuric acid and hydrochloric acid were the analytical reagents, from Chongqing Chuandong Chemical Co., Ltd. Chongqing, China.

2.2. Procedure

As shown in Figure 1, the gangue was crushed and passed through a 160-mesh sieve, then calcined at a high temperature for a certain time to obtain the calcined powder; then, the calcined powder was acid-leached by sulfuric acid or hydrochloric acid at the solid:liquid ratio of 1:3.5, at a temperature of 93 °C, and the stirring speed of 1000 r/min for 1–7 h. After the acid leaching had completed, suction filtration, washing, and drying were carried out. According to the quantity of the filter residue and its chemical components, the extraction ratios of aluminum and iron ions were calculated by the following formulas:
yield of residue = (m2 ÷ m1) × 100%
x = [(m1 × w1% − m2×w2%) ÷ m1w1%] × 100%
where “m1” is the quantity of the calcined powder before leaching, “m2” is the quantity of the filter residue, “w1%” is the mass fraction of aluminum or iron ions in the calcined gangue powder, and “w2%” is the mass fraction of aluminum or iron ions in the filter residue. The compositions of calcined powder before and after acid leaching were detected by XRF, and the concentration of filtrate was calculated accordingly. Then, the USCM was adopted to analyze the leaching kinetics of aluminum and iron ions, and the apparent activation energies were calculated.

2.3. Instrumentation and Characterization

A 6100 type X-ray diffraction instrument (XRD, Shimadzu Company, Japan) was used for CuKα (λ for Kα = 1.54059 Å), 2θ = 3°–65°, with a step width of 0.02°. The primary chemical elemental components were determined by a Supermini200 type X-ray fluorescence spectrometer (XRF, Rigaku Company, Tokyo, Japan). The morphology of the products was identified by a Zeiss EVO18-type scanning electron microscope (SEM, Jena, Germany). The thermal–chemical properties of coal gangue were characterized by a SDT−Q600 type thermal analyzer (TG-DSC, TA Company, Boston, MA, USA), and analyses were performed under the air atmosphere at a constant flow rate of 100 mL/min, where the sensitivity of the microbalance was 0.1 μg, the accuracy of temperature measurement was 0.1 °C, and the heating rate was 20 °C/min.

3. Results and Discussion

3.1. Chemical Components of the Coal Gangue

As shown in Table 1, the contents of SiO2, Al2O3, and Fe2O3 in the gangue accounted for 75.55% of the total quantity. Compared with the coal gangue in other parts of China, the content of the iron element in the sample is higher, whereas the content of the aluminum element is lower.

3.2. XRD Analysis of the Coal Gangue

As seen from Figure 2, the coal gangue is a multiphase mixture [29], and the main phases in the coal gangue are kaolinite, quartz, brookite, montmorillonite, pyrite, and siderite [30], in which kaolinite and quartz are the main components. As shown in Table 1, the iron content is high, but the peaks of the iron-containing materials are lower, which indicates that the crystallinity of the iron-containing materials is low.

3.3. Morphology Analysis of the Coal Gangue

It can be seen from Figure 3 that the coal gangue powders are spherical with an uneven particle size. The maximum particle size is more than 30 μm, whereas the minimum is about 1 μm.

3.4. TG-DSC Analyses of the Coal Gangue

In order to obtain a suitable thermal activation temperature range, the thermal analyses of coal gangue were carried out by TG-DSC. The results are shown in Figure 4.
In Figure 4, the endothermic valley between 100 and 400 °C is mainly caused by the dehydration of water in coal gangue, but the quantity loss is not obvious within this temperature range, which indicates that the content of crystal water in the coal gangue is low. The exothermic peak near 598 °C is mainly caused by the combustion exothermic of coal, pyrite, and other components. When the temperature is higher than 500 °C, kaolinite begins to decompose into metakaolin [31,32], which is an endothermic reaction. The equation is as follows:
Al2Si2O5(OH)4 (kaolinite) → Al2O3·2SiO2 (metakaolin) + 2H2O ↑
The combustion of pyrite is an exothermic reaction, and the reaction is as follows:
4FeS2 (pyrite) + 11O2 → 2Fe2O3 + 8SO2
Overall, the total heat release is larger than the heat absorption, which exhibits an exothermic peak. On the TG curve, the quantity decreases significantly for the combustion of coal and pyrite which produce CO2, SO2, H2O, and other volatile substance products [33]. Meanwhile, siderite is decomposed into FexOy, CO2, or CO, and the volatilizations lead to the quantity loss. When the temperature exceeds 650 °C, kaolin, pyrite, siderite, and other substances have basically produced decomposition, so the quantity is almost unchanged. The low exothermic peak near 800 °C is formed by the continuous combustion and heat release of the residual coal in the coal gangue [34]. The endothermic valley between 900 and 1000 °C is attributed to metakaolin, which is endothermically decomposed into spinel and amorphous silica [35,36]. The main reaction is as follows:
2Al2O3·SiO2 (metakaolin) → Al4Si3O12 (spinel) + SiO2 (amorphous silica)

3.5. Phase Analyses of the Calcined Powders at Different Temperatures

As shown in Figure 5, kaolinite and other minerals can be decomposed completely at 600–800 °C. Therefore, it is conducive to the activation of minerals and the combustion of coal gangue components when enhancing the temperature appropriately. However, too high a temperature will not only waste energy, but also lead to the transformations of metakaolin and other substances into other phases.
There are still crystal inert substances such as quartz and brookite, and the main contents of these substances are silicon and titanium. The gangue rich in aluminum and iron has been transformed into a highly active amorphous state [37], including amorphous Al2O3 and SiO2. Therefore, coal gangue can be activated in the range of 625–750 °C.

3.6. Analyses of the Leaching Effects with Different Acids

Two aliquots with masses of 10 g coal gangue powders were calcined at 750 °C for 2 h, and then the powders were leached by 3.0 mol/L sulfuric acid and 6.0 mol/L hydrochloric acid, respectively. The solid:liquid ratio was 1:3.5, and the whole process was stirred at 93 °C. After the acid leaching reactions, the solid:liquid mixtures were filtered, washed, and dried. After that, the quantities of the filter residues were weighed to obtain the yields of the residues. The results are presented in Figure 6.
As seen from Figure 6, the yields of residue decrease with the extension of time. The yields of residue by sulfuric acid leaching are obviously higher than that of hydrochloric acid.
Hydrochloric acid is a common industrial waste acid, although its demand is not as high as sulfuric acid; therefore, using hydrochloric acid could provide a certain reference for industrial production. In addition, coal gangue contains calcium ions, and adding sulfuric acid can easily produce CaSO4. The equation is as follows:
CaCO3 + H2SO4 = CaSO4 + H2O + CO2
CaSO4 is a substance which is slightly soluble in water. It is easy to cover the surface of the particles and affect the diffusion of ions, which is not conducive to metal ion leaching.
The reaction of hydrochloric acid with CaCO3 produces CaCl2, which is soluble in water and conducive to the dissolution of other ions. The equation is as follows:
CaCO3 + 2HCl = CaCl2 + H2O + CO2
In fact, hydrochloric acid is used very actively in the field of hydrometallurgy, which is used to leach iron or aluminum from coal ash [38], fly ash [39], bauxite [40,41], red mud [42], and coal gangue [43,44]. As can be seen from the results in Figure 6, the effect of hydrochloric acid is significantly better than that of sulfuric acid, so hydrochloric acid was chosen, and the time can be 3–7 h. In order to leach the gangue fully and save energy, the leaching time was determined to be 4 h.
The coal gangue was calcined at 750 °C for 2 h, and after leaching, the phase analyses of the two acid-leached residues are shown in Figure 7.
As shown in Figure 7, the diffraction peaks of the intensity of the quartz and brookite in the filter residue of the leached hydrochloric acid are higher than those of the sulfuric acid, and because more aluminum and iron ions are dissolved by hydrochloric acid, the higher the contents of the remaining quartz and brookite.

3.7. Effects of Calcination Temperature on the Yield of Residue

After the reactions, the reduction quantity was calculated. The results are shown in Figure 8. When the calcination temperature was between 550 and 700 °C, the yield tended to be stable, and the minimum yield was at 675 °C. However, when the calcination temperature was raised to 700–750 °C, the yield increased, indicating that the calcined temperature affects the activity of aluminum and iron ions. In order to activate the aluminum and iron components, as well as remove carbon as much as possible, 675 °C was selected as the proper calcination temperature.

3.8. Effect of Calcination Time on the Yield of Residue

The coal gangue powders were calcined at 675 °C for 30 min, 45 min, 60 min, 75 min, 90 min, 105 min, 120 min, and 135 min. The other conditions are the same as above, and the results are shown in Figure 9.
As shown in Figure 9, when the calcination time was 30–60 min, the yield of residue decreased with the extension of time because the kaolinite, siderite, pyrite, etc., in the coal gangue underwent chemical reactions to produce a large amount of amorphous Al2O3, SiO2, and FexOy, which improved the activity of the calcined powder, thereby decreasing the yield. When the time was extended more than 60 min, the activity of the calcined powder decreased, which led to the increased yield. Therefore, the calcination time was determined to be 1 h. The components of the acid-leached residue are displayed in Table 2.
As seen from Table 2, the primary acid-leached filter residue contained 76.95% of SiO2 and 10.37% of Al2O3. According to the calculations, 91% of iron ions and 66% of aluminum ions have been leached, and the concentration of AlCl3 is 94.26 g/L and that of FeCl3 is 73.13 g/L. The mixed liquid containing aluminum and iron can be used to prepare composite coagulants, or to prepare alumina and iron oxide by a step precipitation method. The residue is mainly silicon- and titanium-containing material, which can be used to prepare other products such as zeolites [45], silica, TiO2, etc.

3.9. Morphology Analysis of Filter Residue

As shown in Figure 10, the particle sizes of the acid-leached filter residues are smaller than that of the ore coal gangue powder, and the largest particle size is about 20 μm. It can also be seen that due to the leaching of iron and aluminum ions, the residue particles become loose, showing flakes and crumbs.

3.10. Analysis of Leaching Kinetics

The components of coal gangue are very complex, in which iron oxide comes from FeCO3 and pyrite, as well as other amorphous iron compounds; aluminum primarily comes from amorphous alumina produced by the decomposition of kaolinite. In the leaching process, acid molecules diffuse into the voids or cracks of quartz or amorphous SiO2, and then react with aluminum and iron oxides. With the reactions, products come out from the residual solid layers, and the reaction interface continues to shrink to the core of the mineral particles. Simultaneously, the residual solid layer continues to thicken, and some solid layers may form debris and peel off under stirring. Therefore, the “USCM” is used to describe the leaching kinetics of aluminum and iron ions [46,47,48].
The reaction model of the leaching process is shown in Figure 11.
Assuming that the calcined coal gangue powder particles are approximately spherical, and the leaching process is controlled by the “solid film diffusion control”, the leaching kinetic equation is as follows:
1 − 2/3x − (1 − x)2/3 = k1t + b1
If the control step is the “interface chemical reaction control”, the equation is as follows:
1 − (1 − x)1/3 = k2t + b2
When the leaching process is jointly controlled by the “solid film diffusion control” and the “interface chemical reaction control” together, it is called the “mixing control”, and the equation is [49]:
(1 − x)−1/3 − 1 + 1/3ln(1 − x) = k3t + b3
where “k1” is the solid–liquid diffusion rate constant, “k2” is the rate constant of interfacial chemical reaction, “k3” is the rate constant of mixing control reaction, “x” is the leaching ratio, “t” is the leaching time, and “b1, b2 and b3” are the constant terms.
In order to determine the control type of hydrochloric acid leaching reactions, the extraction ratios of aluminum and iron ions changing with temperature and time were assessed. The temperature increased from 303.15 K to 363.15 K with a step of 20 K, the time was from 10 min to 5 h, the H+ concentration is 6 mol/L, and the stirring speed was 1000 r/min. Since the contents of calcium and magnesium ions in coal the gangue were very low (Cao 0.05%, MgO 0.22%), and TiO2 (6.22%) did not react with dilute hydrochloric acid, we ignored the effects of calcium, magnesium, and titanium ions on leaching. The results are shown in Figure 12 and Figure 13.
A trial method is used to substitute the obtained results into the kinetic equations of the different control steps, plotting 1 − 2/3x − (1 − x)2/3, 1 − (1 − x)1/3 and (1 − x)−1/3 − 1 + 1/3ln(1 − x) versus t, respectively, and then using these three models to fit the obtained data linearly. The results of aluminum ions are shown in Figure 14, Figure 15 and Figure 16, and so the iron ions are shown in Figure 17, Figure 18 and Figure 19.
As can be seen from the fitting equations and the R2 values of these three mathematical models, whether aluminum ion leaching or iron ion leaching, the R2 values of the mixing control equation at 30 °C, 50 °C, 70 °C, and 90 °C were all better than that of single “solid film diffusion control” or “interface chemical reaction control”, and were all closer to 1. Therefore, hydrochloric acid leaching aluminum and iron ions kinetic equations from the calcined coal gangue powder are all accord with the “mixing control” models.

3.11. Calculation of the Apparent Activation Energy

To obtain the activation energies of aluminum and iron ions extraction under hydrochloric acid leaching conditions, the Arrhenius formula was used to determine the apparent activation energies of the leaching experiments. According to the Arrhenius formula:
k = AeEa/RT
lnk = −Ea/RT + lnA
where “k” is the apparent reaction rate constant, “Ea” is the apparent activation energy, kJ/mol, “R” is the molar gas constant, 8.314 J/(mol K), and “A” is a constant. The fitting curves of lnk−1/T of aluminum and iron ions are shown in Figure 20 and Figure 21. The apparent activation energies can be calculated from the slope of the fitting equations based on Arrhenius formula, where the slope is −Ea/RT and the intercept is lnA.
As seen from Figure 20, the correlation fitting degree of the lnk−1/T curve (R2 = 0.95901) obtained from the “mixing control” fitting is significantly better than that of the diffusion of solid film control (R2 = 0.91902) and interfacial chemical reaction control (R2 = 0.79103). Similarly, the correlation coefficient R2 of the iron ions fitting equation using the “mixing control” model is also better than that of the other two models (Figure 21), which further indicates that the leaching ratios of aluminum and iron ions from calcined coal gangue powder by hydrochloric acid leaching are both affected by the “mixing control”.
According to Figure 20 and Figure 21, we can derive:
lnk(Al) = −6675.7400/T + 14.30627
lnk(Fe) = −6712.34419/T + 15.81136
Ea(Al) = 6675.7400R = 55.5 × 103 J/mol = 55.5 kJ/mol
Ea(Fe) = 6712.3442R = 55.8 × 103 J/mol = 55.8 kJ/mol
The apparent activation energies of aluminum and iron ion leaching are 55.5 kJ/mol and 55.8 kJ/mol, respectively.

4. Conclusions

In the study of extracting aluminum and iron ions from coal gangue, the proper conditions were obtained when the gangue powder was less than 160 mesh, the calcination temperature was 675 °C, the time was 60 min, and the calcined coal gangue was leached by 6 mol/L hydrochloric acid at 93 °C for 4 h, where the leaching ratio of iron ions was 91% and that of aluminum ions was 66%.
The process of leaching aluminum and iron ions from the calcined coal gangue powder with hydrochloric acid conforms to the kinetic model: (1 − x)−1/3 − 1 + 1/3ln(1 − x) = k3t + b3. According to the Arrhenius equation, the apparent activation energies of the leaching reactions of aluminum and iron are 55.5 kJ/mol and 55.8 kJ/mol, respectively. The leaching process of aluminum and iron ions are controlled by the “mixing control”.

Author Contributions

D.K.: Literature search, Conceptualization, Methodology, Investigation, Visualization, Experiment, Data analysis; Writing—Original Draft, Writing—Review and Editing; R.J.: Investigation, Data analysis; Writing—Review and Editing; Z.Z.: Investigation, Writing—Review and Editing; S.S.: Investigation, Data analysis; Writing—Review and Editing; S.F.: Investigation, Writing—Review and Editing; M.L.: Writing—Review and Editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Guizhou Provincial Science and Technology Department: 2018]1142 and [2018]1415, Guizhou Provincial Education Department: [2017]054, Liupanshui City Science and Technology Foundation: 52020-2019-05-17.

Data Availability Statement

The study did not report any data.

Acknowledgments

The study was financially supported by Guizhou Provincial Education Department’s Scientific and Technological Innovation Team Project (NO. [2017]054), the Guizhou Science and Technology Foundation Project (NO. [2018]1142 and [2018]1415), and the Liupanshui City Science and Technology Foundation (NO. 52020-2019-05-17).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ma, H.Q.; Zhu, H.G.; Wu, C.; Chen, H.Y.; Sun, J.W.; Liu, J.Y. Study on compressive strength and durability of alkali-activated coal gangue-slag concrete and its mechanism. Powder Technol. 2020, 368, 112–124. [Google Scholar] [CrossRef]
  2. Huang, G.D.; Ji, Y.S.; Li, J.; Hou, Z.H.; Dong, Z.C. Improving strength of calcinated coal gangue geopolymer mortars via increasing calcium content. Construct. Build. Mater. 2018, 166, 760–768. [Google Scholar] [CrossRef]
  3. Yang, F.L.; Yang, Y.; Li, G.K.; Sang, N. Heavy metals in soil from gangue stacking areas increases children health risk and causes developmental neurotoxicity in zebrafish larvae. Sci. Total Environ. 2021, 794, 148629. [Google Scholar] [CrossRef] [PubMed]
  4. Gomo, M.; Vermeulen, D. Hydrogeochemical characteristics of a flooded underground coal mine groundwater system. J. Afr. Earth. Sci. 2014, 92, 68–75. [Google Scholar] [CrossRef]
  5. Li, M.; Zhang, J.X.; Li, A.L.; Zhou, N. Reutilisation of coal gangue and fly ash as underground back fill materials for surface subsidence control. J. Clean. Prod. 2020, 254, 120113. [Google Scholar] [CrossRef]
  6. Xu, H.L.; Song, W.J.; Cao, W.B.; Shao, G.; Lu, H.X.; Yang, D.Y.; Chen, D.; Zhang, R. Utilization of coal gangue for the production of brick. J. Mater. Cycles Waste Manag. 2017, 19, 1270–1278. [Google Scholar] [CrossRef]
  7. Xu, B.H.; Liu, Q.F.; Ai, B.; Ding, S.L.; Frost, R.L. Thermal decomposition of selected coal gangue. J. Therm. Anal. Calorim. 2018, 131, 1–10. [Google Scholar] [CrossRef]
  8. Guan, J.; Lu, M.; Yao, X.H.; Wang, Q.; Wang, D.C.; Yang, B.; Liu, H. An experimental study of the road performance of cement stabilized coal gangue. Crystals 2021, 11, 993. [Google Scholar] [CrossRef]
  9. Zhou, M.; Dou, Y.W.; Zhang, Y.Z.; Zhang, Y.Q.; Zhang, B.Q. Effects of the variety and content of coal gangue coarse aggregate on the mechanical properties of concrete. Construct. Build. Mater. 2019, 220, 386–395. [Google Scholar] [CrossRef]
  10. Yang, J.; Zhang, X.; Liu, Y.Q.; Ma, H.W.; Feng, W.W.; Zeng, C. Synthesis of ultrafine silica powder using the silica source alkali-leached from coal gangue. Integr. Ferroelectr. 2015, 161, 10–17. [Google Scholar] [CrossRef]
  11. Wu, R.D.; Dai, S.B.; Jian, S.W.; Huang, J.; Tan, H.B.; Li, B.D. Utilization of solid waste high-volume calcium coal gangue in autoclaved aerated concrete: Physico-mechanical properties, hydration products and economic costs. J. Clean. Prod. 2021, 278, 123416. [Google Scholar] [CrossRef]
  12. Bu, N.J.; Liu, X.M.; Song, S.L.; Liu, J.H.; Yang, Q.; Li, R.; Zheng, F.; Yan, L.; Zhen, Q.; Zhang, J. Synthesis of NaY zeolite from coal gangue and its characterization for lead removal from aqueous solution. Adv. Powder Technol. 2020, 31, 2699–2710. [Google Scholar] [CrossRef]
  13. Gao, Y.; Huang, J.D.; Li, M.; Dai, Z.R.; Jiang, R.L.; Zhang, J.X. Chemical modification of combusted coal gangue for U(VI) adsorption: Towards a waste control by waste strategy. Sustainability 2021, 13, 8421. [Google Scholar] [CrossRef]
  14. Han, L.; Ren, W.; Wang, B.; He, X.; Ma, L.; Huo, Q.; Wang, J.; Bao, W.; Chang, L. Extraction of SiO2 and Al2O3 from coal gangue activated by supercritical water. Fuel 2019, 253, 1184–1192. [Google Scholar] [CrossRef]
  15. Tang, Z.H. Preparation of nanometer Al(OH)3 powder using coal gangue by polyacrylamide dispersant-carbonization method. Adv. Mater. Res. 2011, 160–162, 307–313. [Google Scholar] [CrossRef]
  16. Dong, L.; Liang, X.; Song, Q.; Gao, G.; Song, L.; Shu, Y.; Shu, X. Study on Al2O3 extraction from activated coal gangue under different calcination atmospheres. J. Therm. Sci. 2017, 26, 570–576. [Google Scholar] [CrossRef]
  17. Guo, Y.X.; Zhao, Q.; Yan, K.Z.; Cheng, F.Q.; Lou, H.H. Novel process for alumina extraction via the coupling treatment of coal gangue and bauxite red mud. Ind. Eng. Chem. Res. 2014, 53, 4518–4521. [Google Scholar] [CrossRef]
  18. Guo, Y.X.; Lv, H.B.; Yang, X.; Cheng, F.Q. AlCl3·6H2O recovery from the acid leaching liquor of coal gangue by using concentrated hydrochloric inpouring. Sep. Purif. Technol. 2015, 151, 177–183. [Google Scholar] [CrossRef]
  19. Cheng, Y.; Hongqiang, M.; Hongyu, C.; Jiaxin, W.; Jing, S.; Zonghui, L.; Mingkai, Y. Preparation and characterization of coal gangue geopolymers. Construct. Build. Mater. 2018, 187, 318–326. [Google Scholar] [CrossRef]
  20. Wang, R.G.; Yun, L.X. Preparation and wastewater treatment of polymeric aluminum chloride from coal gangue. Adv. Mater. Res. 2012, 518–523, 780–783. [Google Scholar] [CrossRef]
  21. Zhang, S.; Zhang, N.; Zhao, W.; Lan, D.; Hao, G.; Yi, X.; Yang, Z.; Hu, J.; He, W.; Liu, Y.; et al. Green preparation of hierarchical porous C/SiOx composites from coal gangue as anodes for Li-ion batteries. Solid State Ionics 2021, 371, 115772. [Google Scholar] [CrossRef]
  22. Zhou, L.; Zhou, H.J.; Hu, Y.X.; Yan, S.; Yang, J.L. Adsorption removal of cationic dyes from aqueous solutions using ceramic adsorbents prepared from industrial waste coal gangue. J. Environ. Manag. 2019, 234, 245–252. [Google Scholar] [CrossRef] [PubMed]
  23. Li, J.F.; Li, J.G.; Liu, X.Y.; Du, Z.P.; Cheng, F.Q. Effect of silicon content on preparation and coagulation performance of poly-silicic-metal coagulants derived from coal gangue for coking wastewater treatment. Sep. Purif. Technol. 2018, 202, 149–156. [Google Scholar] [CrossRef]
  24. Luo, K.J.; Ren, G.M. Coagulant prepared by gangue and its application in treatment of coal washing wastewater. Adv. Mater. Res. 2010, 142, 270–273. [Google Scholar] [CrossRef]
  25. Xiao, J.; Li, F.; Zhong, Q.; Bao, H.; Wang, B.; Huang, J.; Zhang, Y. Separation of aluminum and silica from coal gangue by elevated temperature acid leaching for the preparation of alumina and SiC. Hydrometallurgy 2015, 155, 118–124. [Google Scholar] [CrossRef]
  26. Yang, Q.; Zhang, F.; Deng, X.; Guo, H.; Zhang, C.; Shi, C.; Zeng, M. Extraction of alumina from alumina rich coal gangue by a hydro-chemical process. Roy. Soc. Open Sci. 2020, 7, 192132. [Google Scholar] [CrossRef]
  27. Johnston, C.J.; Pepper, R.A.; Martens, W.N.; Couperthwaite, S. Improvement of aluminium extraction from low-grade kaolinite by iron oxide impurities: Role of clay chemistry and morphology. Miner. Eng. 2022, 176, 107346. [Google Scholar] [CrossRef]
  28. Remya, P.N.; Kazantzis, N.K.; Emmert, M.H. Selective process steps for the recovery of scandium from jamaican bauxite residue (red mud). ACS Sustain. Chem. Eng. 2018, 6, 1478–1488. [Google Scholar]
  29. Ma, H.Q.; Yi, C.; Zhu, H.G.; Dong, Z.C.; Chen, H.Y.; Wang, J.X.; Li, D.Y. Compressive strength and durability of coal gangue aggregate concrete. Mater. Rep. 2018, 32, 2390–2395. [Google Scholar]
  30. Zhang, M.S.; Xiao, L.W. Preparation of a gangue-based X-type zeolite molecular sieve as a multiphase fenton catalyst and its catalytic performance. ACS Omega 2021, 6, 18414–18425. [Google Scholar] [CrossRef]
  31. Zhang, Y.Y.; Zhang, Z.Z.; Zhu, M.M.; Cheng, F.Q.; Zhang, D.K. Decomposition of key minerals in coal gangues during combustion in O2/N2 and O2/CO2 atmospheres. Appl. Therm. Eng. 2019, 148, 977–983. [Google Scholar] [CrossRef]
  32. Gasparini, E.; Tarantino, S.; Ghigna, P.; Riccardi, M.P.; Cedillo-González, E.I.; Siligardi, C.; Zema, M. Thermal dehydroxylation of kaolinite under isothermal conditions. Appl. Clay Sci. 2013, 80–81, 417–425. [Google Scholar] [CrossRef]
  33. Zhang, Y.Y.; Ge, X.L.; Nakano, J.; Liu, L.L.; Wang, X.D.; Zhang, Z.T. Pyrite transformation and sulfur dioxide release during calcination of coal gangue. RSC Adv. 2014, 4, 42506–42513. [Google Scholar] [CrossRef]
  34. Xie, M.Z.; Liu, F.Q.; Zhao, H.L.; Ke, C.Y.; Xu, Z.Q. Mineral phase transformation in coal gangue by high temperature calcination and high-efficiency separation of alumina and silica minerals. J. Mater. Res. Technol. 2021, 14, 2281–2288. [Google Scholar] [CrossRef]
  35. Ptáček, P.; Šoukal, F.; Opravil, T.; Nosková, M.; Havlica, J.; Brandštetr, J. The kinetics of Al-Si spinel phase crystallization from calcined kaolin. J. Solid State Chem. 2010, 183, 2565–2569. [Google Scholar] [CrossRef]
  36. Kenne Diffo, B.B.; Elimbi, A.; Cyr, M.; Dika Manga, J.; Tchakoute Kouamo, H. Effect of the rate of calcination of kaolin on the properties of metakaolin-based geopolymers. J. Asian Ceram. Soc. 2015, 42, 130–138. [Google Scholar] [CrossRef] [Green Version]
  37. Cao, Z.; Cao, Y.D.; Dong, H.J.; Zhang, J.S.; Sun, C.B. Effect of calcination condition on the microstructure and pozzolanic activity of calcined coal gangue. Int. J. Miner. Process. 2016, 146, 23–28. [Google Scholar] [CrossRef]
  38. Valeev, D.; Kunilova, I.; Shoppert, A.; Salazar-Concha, C.; Kondratiev, A. High-pressure HCl.leaching of coal ash to extract Al into a chloride solution with further use as a coagulant for water treatment. J. Clean. Prod. 2020, 276, 123206. [Google Scholar] [CrossRef]
  39. Valeev, D.; Mikhailova, A.; Atmadzhidi, A. Kinetics of iron extraction from coal fly ash by hydrochloric acid leaching. Metals 2018, 8, 533. [Google Scholar] [CrossRef] [Green Version]
  40. Valeev, D.V.; Lainer, Y.A.; Mikhailova, A.B.; Dorofievich, I.V.; Zheleznyi, M.V.; Gol’dberg, M.A. Reaction of bauxite with hydrochloric acid under autoclave conditions. Metallurgist 2016, 60, 204–211. [Google Scholar] [CrossRef]
  41. Valeev, D.V.; Lainer, Y.A.; Pak, V.I. Autoclave leaching of boehmite-kaolinite bauxites by hydrochloric acid. Inorg. Mater. Appl. Res. 2016, 7, 272–277. [Google Scholar] [CrossRef]
  42. Zinoveev, D.; Grudinsky, P.; Zhiltsova, E.; Grigoreva, D.; Volkov, A.; Dyubanov, V.; Petelin, A. Research on high-pressure hydrochloric acid leaching of scandium, aluminum and other valuable components from the non-magnetic tailings obtained from red mud after iron removal. Metals 2021, 11, 469. [Google Scholar] [CrossRef]
  43. Cheng, F.; Cui, L.; Miller, J.D.; Wang, X. Aluminum leaching from calcined coal waste using hydrochloric acid solution. Miner. Procoss. Extr. Metall. Rev. 2012, 33, 391–403. [Google Scholar] [CrossRef]
  44. Zhang, L.; Wang, H.; Li, Y. Research on the extract Al2O3 from coal gangue. Adv. Mater. Res. 2012, 524–527, 1947–1950. [Google Scholar] [CrossRef]
  45. Kong, D.S.; Jiang, R.L. Preparation of NaA zeolite from high iron and quartz contents coal gangue by acid leaching-alkali melting activation and hydrothermal synthesis. Crystals 2021, 11, 1198. [Google Scholar] [CrossRef]
  46. Gui, Q.H.; Khan, M.I.; Wang, S.X.; Zhang, L.B. The ultrasound leaching kinetics of gold in the thiosulfate leaching process catalysed by cobalt ammonia. Hydrometallurgy 2020, 196, 105426. [Google Scholar] [CrossRef]
  47. Dickinson, C.F.; Heal, G.R. Solid-liquid diffusion controlled rate equations. Thermochim. Acta 1999, 340, 89–103. [Google Scholar] [CrossRef]
  48. Rao, S.; Yang, T.; Zhang, D.; Liu, W.; Chen, L.; Hao, Z.; Xiao, Q.; Wen, J. Leaching of low grade zinc oxide ores in NH4Cl-NH3 solutions with nitrilotriacetic acid as complexing agents. Hydrometallurgy 2015, 158, 101–106. [Google Scholar] [CrossRef]
  49. Valeev, D.; Pankratov, D.; Shoppert, A.; Sokolov, A.; Kasikov, A.; Mikhailova, A.; Salazar-Concha, C.; Rodionov, I. Mechanism and kinetics of iron extraction from high silica boehmite–kaolinite bauxite by hydrochloric acid leaching. Trans. Nonferrous Met. Soc. China 2021, 31, 3128–3149. [Google Scholar] [CrossRef]
Figure 1. Flow chart of the experiments.
Figure 1. Flow chart of the experiments.
Minerals 12 00215 g001
Figure 2. XRD spectrum of coal gangue powder.
Figure 2. XRD spectrum of coal gangue powder.
Minerals 12 00215 g002
Figure 3. SEM image of raw gangue powders.
Figure 3. SEM image of raw gangue powders.
Minerals 12 00215 g003
Figure 4. TG – DSC chart of the coal gangue.
Figure 4. TG – DSC chart of the coal gangue.
Minerals 12 00215 g004
Figure 5. XRD spectra of the calcined powders at different temperatures. H: hematite, Fe2O3.
Figure 5. XRD spectra of the calcined powders at different temperatures. H: hematite, Fe2O3.
Minerals 12 00215 g005
Figure 6. The effect of acid leaching time on the yield of residue.
Figure 6. The effect of acid leaching time on the yield of residue.
Minerals 12 00215 g006
Figure 7. XRD spectra of the filter residue after acid leaching with different acids for 6 h.
Figure 7. XRD spectra of the filter residue after acid leaching with different acids for 6 h.
Minerals 12 00215 g007
Figure 8. Effects of calcination temperature on the yield of residue.
Figure 8. Effects of calcination temperature on the yield of residue.
Minerals 12 00215 g008
Figure 9. Effect of different calcination time on the yield of residue at 675 °C.
Figure 9. Effect of different calcination time on the yield of residue at 675 °C.
Minerals 12 00215 g009
Figure 10. SEM image of the residue.
Figure 10. SEM image of the residue.
Minerals 12 00215 g010
Figure 11. Schematic diagram of acid leaching process.
Figure 11. Schematic diagram of acid leaching process.
Minerals 12 00215 g011
Figure 12. Effects of temperature on the extraction ratios of aluminum ions.
Figure 12. Effects of temperature on the extraction ratios of aluminum ions.
Minerals 12 00215 g012
Figure 13. Effects of temperature on the extraction ratios of iron ions.
Figure 13. Effects of temperature on the extraction ratios of iron ions.
Minerals 12 00215 g013
Figure 14. Relationships between 1 − 2/3x − (1 − x)2/3 and time of aluminum ions.
Figure 14. Relationships between 1 − 2/3x − (1 − x)2/3 and time of aluminum ions.
Minerals 12 00215 g014
Figure 15. Relationships between 1 − (1 − x)1/3 and time of aluminum ions.
Figure 15. Relationships between 1 − (1 − x)1/3 and time of aluminum ions.
Minerals 12 00215 g015
Figure 16. Relationships between (1 − x)−1/3 − 1 + 1/3ln(1 − x) and time of aluminum ions.
Figure 16. Relationships between (1 − x)−1/3 − 1 + 1/3ln(1 − x) and time of aluminum ions.
Minerals 12 00215 g016
Figure 17. Relationships between 1 − 2/3x − (1 − x)2/3 and time of iron ions.
Figure 17. Relationships between 1 − 2/3x − (1 − x)2/3 and time of iron ions.
Minerals 12 00215 g017
Figure 18. Relationships between 1 − (1 − x)1/3 and time of iron ions.
Figure 18. Relationships between 1 − (1 − x)1/3 and time of iron ions.
Minerals 12 00215 g018
Figure 19. Relationships between (1 − x)−1/3 − 1 + 1/3ln(1 − x) and time of iron ions.
Figure 19. Relationships between (1 − x)−1/3 − 1 + 1/3ln(1 − x) and time of iron ions.
Minerals 12 00215 g019
Figure 20. Curves of lnK−1/T to aluminum ions.
Figure 20. Curves of lnK−1/T to aluminum ions.
Minerals 12 00215 g020
Figure 21. Curves of lnK−1/T to iron ions.
Figure 21. Curves of lnK−1/T to iron ions.
Minerals 12 00215 g021
Table 1. Main components of the coal gangue.
Table 1. Main components of the coal gangue.
Componentswt/%
SiO242.18
Al2O320.43
Fe2O311.94
CaO2.95
MgO1.61
MnO0.25
P2O50.30
TiO23.77
S0.54
K2O1.24
Na2O0.40
FC and Loss14.39
Table 2. Main components of the residue.
Table 2. Main components of the residue.
Componentswt/%
SiO276.95
Al2O310.37
Fe2O31.00
CaO0.05
MgO0.22
TiO26.22
MnO0.04
P2O50.06
S0.01
K2O1.01
Na2O0.38
FC and others3.69
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kong, D.; Zhou, Z.; Jiang, R.; Song, S.; Feng, S.; Lian, M. Extraction of Aluminum and Iron Ions from Coal Gangue by Acid Leaching and Kinetic Analyses. Minerals 2022, 12, 215. https://doi.org/10.3390/min12020215

AMA Style

Kong D, Zhou Z, Jiang R, Song S, Feng S, Lian M. Extraction of Aluminum and Iron Ions from Coal Gangue by Acid Leaching and Kinetic Analyses. Minerals. 2022; 12(2):215. https://doi.org/10.3390/min12020215

Chicago/Turabian Style

Kong, Deshun, Zihan Zhou, Rongli Jiang, Shuojiang Song, Shan Feng, and Minglei Lian. 2022. "Extraction of Aluminum and Iron Ions from Coal Gangue by Acid Leaching and Kinetic Analyses" Minerals 12, no. 2: 215. https://doi.org/10.3390/min12020215

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