Next Article in Journal
Hydrogenetic and Diagenetic Controls on the Specific Surface Area of Polymetallic Nodules in Deep Ocean Basins
Next Article in Special Issue
Influence of Sulfate and Nitrate for Lanthanum (III) Adsorption on Bentonite: Implications for Rare Earth Wastewater Disposal
Previous Article in Journal
Numerical Modeling of the Hydrothermal Metallogenic Mechanism Associated with the Ergu Pb-Zn Deposit, Heilongjiang, China: An Example of Pore-Fluid Convection Controlled Mineralization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Natural Pyrite as a Catalyst for a Fenton Reaction to Enhance Xanthate Degradation in Flotation Tailings Wastewater

1
Faculty of Land Resource Engineering, Kunming University of Science and Technology, Kunming 650093, China
2
State Key Laboratory of Complex Nonferrous Metal Resources Clean Utilization, Kunming University of Science and Technology, Kunming 650093, China
3
Yunnan Key Laboratory of Green Separation and Enrichment of Strategic Mineral Resources, Kunming 650093, China
*
Author to whom correspondence should be addressed.
Minerals 2023, 13(11), 1429; https://doi.org/10.3390/min13111429
Submission received: 20 July 2023 / Revised: 31 October 2023 / Accepted: 7 November 2023 / Published: 10 November 2023
(This article belongs to the Special Issue Comprehensive Utilization of Mineral Processing Wastewater)

Abstract

:
The efficient treatment of mineral-processing wastewater has attracted soaring interest recently. This study’s objective was to degrade xanthate from flotation tailings wastewater using a pyrite-catalyzed Fenton system. A sodium butyl xanthate (SBX) removal rate of more than 96% was achieved via the method under optimal conditions (a H2O2 concentration of 0.5 mM, a FeS2 concentration of 0.5 g/L, an initial SBX concentration of 100 mg/L, and a natural pH of 9.36 ± 0.5), which is 12.85% higher than with a H2O2 system. An appropriate concentration of natural pyrite can act as a catalyst to significantly improve the oxidation capacity of H2O2. Additionally, the results of electron paramagnetic resonance and quenching measurements suggest that hydroxyl radicals (•OH) are the main active species in the H2O2-FeS2 system. The possible reaction mechanism is proposed. The H2O2 adsorbs onto the pyrite surfaces and reacts with Fe2+, triggering the formation of •OH and Fe3+. The •OH most likely attacks the SBX that adsorbs on the pyrite surface or exists in the solution and promotes the transformation of the SBX anion (C4H9OCS2) into the intermediate butyl xanthate peroxide (BPX, C4H9OCS2O). Finally, BPX intermediates are likely further oxidized to smaller products such as SO42−, CO2, and H2O under the ongoing attack of •OH.

1. Introduction

Froth flotation, on the basis of the different surface properties of minerals, has become the most common and effective beneficiation method for the separation, purification, and enrichment of valuable components in ore [1,2]. Xanthate is one of the most common organic-sulfur-based collectors for the froth flotation of sulfide ores as well as a multitude of oxidized ores [3,4]. It has been reported that the consumption of alkyl xanthate is expected to reach approximately 371,826,000 tons by 2025 [5]. However, the flotation circuit only consumes half of the alkyl xanthate amount, and the other half is discharged to tailing ponds [6]. It is well recognized that xanthate residue in flotation tailings wastewater can not only seriously pollute the ecological environment around the mining area, but it can also damage the human nervous system and liver system [7,8]. Furthermore, when recycling beneficiation wastewater, this remaining agent will affect the flotation index and cause adverse effects on the flotation selectivity of the minerals [9]. Thus, it is of great significance to effectively treat the residual xanthate in flotation wastewater for the sustainable development of the mineral-processing industry and for environmental protection.
Recently, environmental issues related to the mineral-processing industry have received increased attention from researchers. The efficient treatment of mineral-processing wastewater is the focus of research in this field. Traditional processes have been developed to remove xanthate from flotation wastewater, including coagulation–flocculation [10], adsorption [11,12], biodegradation [13,14], and chemical oxidation [15]. On the other hand, there are still many issues in the treatment of flotation xanthate wastewater via traditional methods. For example, the methods of coagulation–flocculation and adsorption only transfer xanthate molecules from wastewater to solid products or adsorbents, and can produce secondary pollution [16,17]. The microbial method takes a long time to degrade xanthate wastewater [4]. The conventional chemical oxidation methods have an extremely low treatment efficiency, and the treated xanthate wastewater usually fails to meet emission standards for release into the environment [10]. Therefore, it is vitally important to develop a green and efficient method for solving the problem of flotation tailings wastewater pollution.
During the last few decades, efforts to treat mineral-processing wastewater via advanced oxidation processes (AOPs) have been ongoing [18,19,20]. The Fenton oxidation process, as a green oxidation technology, has attracted significant attention among AOPs due to its strong oxidative capacity for the degradation of organic contaminants [21,22]. The main reaction mechanism of the classic Fenton reaction induced by aqueous Fe2+ is well known to be the decomposition of H2O2 to •OH (2.76 V) [22,23,24,25] having a higher redox potential than ozone (2.07 V), permanganate (1.68 V), and persulfate (2.01 V) [26]. This makes the Fenton reaction more effective and thermodynamically feasible for the oxidative degradation of organic contaminants. Despite its strong oxidation capacity, the classic homogeneous Fenton reaction has some critical disadvantages such as the rapid precipitation of Fe(OH)3, which generates significant amounts of sludge and causes the early termination of the reaction. Thus, a low pH is always required [27]. Recently, heterogeneous iron sources such as zerovalent iron (ZVI) and Fe-containing minerals (ferrihydrite, hematite, goethite, lepidocrocite, magnetite, and pyrite) playing roles of surface catalysts have been used to produce hydroxyl radicals during the Fenton reaction instead of an aqueous iron source [28,29,30,31]. With advantages over classic Fenton reactions such as less iron sludge formation, a wider working pH range, and the possible recycling of the iron promoter, Fenton systems catalyzed by heterogeneous Fe-containing solid catalysts have received soaring attention [32,33,34].
Pyrite (FeS2) is one of the most abundant sulfide minerals on the Earth and contains a high content of iron [35]. Pyrite can be considered as a catalyst for wastewater treatment due to its low cost and environmentally friendly characteristics. Many researchers have confirmed the cost-effectiveness of pyrite as a catalyst for the removal of water pollutants [36,37,38,39]. Farzaneh Rahimi et al. investigated pyrite nanoparticles derived from mine waste as efficient catalysts for the activation of persulfates for the degradation of tetracycline. The results demonstrated that PMS was activated by using pyrite, and the pyrite/PMS process significantly decreased the nephrotoxicity (90%) and hepatotoxicity (85%) effect of tetracycline [36]. Zhang used an enhanced Fenton system catalyzed by heterogeneous pyrite to degrade nitrobenzene. The results indicated that the appropriate Fe2+ concentration generated by constant dissolution from a pyrite surface and by recycling Fe3+ contributed to this enhancement [37]. Research reported by Hyeongsu Che showed that trichloroethylene (TCE) was oxidatively degraded by •OH in the pyrite Fenton system and its degradation kinetics was significantly enhanced by the catalysis of pyrite to form •OH by decomposing H2O2 [38]. Lian investigated the degradation of tris(2-chloroethyl) phosphate (TCEP) as a category of OPE by pyrite (FeS2)-activated persulfate (PS). The results confirmed that nearly 100% degradation of TCEP was achieved after 120 min in the FeS2-PS system, and Fe3+ and Cl could accelerate the degradation rate of TCEP [39]. Although numerous studies have demonstrated that pyrite could be effectively used as a heterogeneous catalyst for the removal of organic compounds from water, currently, limited basic knowledge is available on the use of natural pyrite as a catalyst for the development of enhanced remediation technologies to treat environmental contaminants, especially xanthate organics in flotation tailings wastewater.
Accordingly, pyrite was selected as a heterogeneous iron source to activate the Fenton reaction for the degradation of xanthate in flotation tailings wastewater. The main goals of this study were as follows: (1) determining the catalytic potential of pyrite in activating H2O2 for the generation of reactive species; (2) evaluating various operational conditions (concentrations of pyrite, H2O2, and SBX); (3) assessing the contribution of free radicals in degradation reactions; (4) proposing a possible degradation pathway. This study may provide a strong basis for the green and sustainable treatment of mineral-processing wastewater.

2. Materials and Methods

2.1. Materials and Reagents

Sodium butyl xanthate (SBX) was supplied by Zhuzhou Flotation Reagents Ltd., Zhuzhou, Hunan, China, with a purity of >98%. H2O2 (non-stabilized, 30 wt%) was purchased from Yunnan Jingrui Technology Co., Ltd., Yunnan, China. 5,5-dimethyl-1-pyrolin-N-oxide (DMPO, >99%) and TBA (tert-butyl alcohol, >99%) were produced by the Aladdin Industrial Corporation, Shanghai, China. The DMPO was serviced as a hydroxyl radical (•OH) trapping agent, and the TBA was used as a hydroxyl radical (•OH) scavenger. Ultra-pure water (resistivity 1.83 × 107 Ω•cm) was acquired from a Milli-Q5O purification system (Billerica, MA, USA). The pure pyrite used in this experiment was mined from Yunnan, China, and was finely ground in an agate mortar to obtain a 200-mesh powder (particle size, <75 μm). The chemical composition of the pyrite samples was analyzed, and the results confirmed that the purity of the pyrite samples was over 95%, of which the Fe content was 45.39% and the S content was 52.58%. Additionally, the pyrite sample was further analyzed via powder X-ray diffraction (XRD) and only the signal from pyrite was observed (Figure 1).

2.2. Degradation Experiments of Sodium Butyl Xanthate (SBX)

For the degradation experiments, 50 mL amounts of the various initial concentrations (50, 75, 100, 125, and 150 mg/L) of SBX solutions were used as simulated flotation tailing wastewater samples and were placed in a 150 mL beaker. The initial pH was 9.36 ± 0.5. After that, certain amounts of H2O2 (if present) and pyrite catalyst (if present) were added into the solution to initiate the degradation reaction. Furthermore, the reaction system was stirred continuously with a magnetic stirrer at 500 rpm to ensure that the SBX solution was well mixed with the reagents and pyrite. Notably, no pH adjustment was performed in the whole process, and a pH meter (pH-25, Rex Electric Chemical) was used to measure the solution’s pH value at the beginning and end of each experiment. All the experiments were carried out at room temperature (25 °C). After a period of reaction (5~85 min), a part of the solution was filtered through a 0.45 μm pore size filter paper. Then, the filtrates were analyzed using an ultraviolet–visible (UV-vis) spectrophotometer. The absorbance of the filtrates was measured at 301 nm to investigate variations in the SBX concentrations during the degradation. The degradation efficiency of SBX was calculated using Equation (1):
η = C 0 C t C 0 × 100 %
where η (%) represents the degradation efficiency, C0 (mg/L) is the initial concentration of SBX, and Ct (mg/L) is the SBX concentration after degradation.

2.3. Analytical Methods

UV full-spectrum scanning (UV-2700, Shimadzu, Japan; scanning range: 200–400 nm) was used to investigate variations in intermediates during SBX degradation. The concentrations of iron ions in the solutions were determined using the spectrophotometric method. Specifically, the standard solution of Fe2+ was prepared with ammonium ferrous sulfate, and the standard curve between Fe2+ concentrations and absorbances was established. Then, 10 mL of the reaction solution was added to a colorimetric tube. A volume of 1 mL of 50 mM hydroxylamine hydrochloride solution was added to the solution, reducing Fe3+ to Fe2+, and the Fe2+ concentration was measured at 510 nm using a ferrozine method on the above-mentioned UV/visible spectrophotometer [40]. Also, the Fe3+ concentrations in the solution could be calculated based on the difference values between the total iron concentration and the Fe2+ concentration [41].

2.4. Capture of Active Species Experiments

The possible mechanism of SBX degradation was initially explored via a free radical capture test and electron paramagnetic resonance (EPR, Bruker, Germany). The EPR test was performed by using DMPO as the trapping agent to identify the generation of hydroxyl radicals (•OH) in the H2O2 solution system and H2O2-FeS2 solution system. Additionally, TBA (200 mg/L) was added to the H2O2-FeS2 system to scavenge hydroxyl radicals for contrast testing with the aim of investigating the role of •OH in the SBX degradation.

3. Results and Discussion

3.1. The Degradation of SBX via the Pyrite-Catalyzed Fenton System

3.1.1. The Effect of H2O2 Concentration

The effect of H2O2 concentration on the degradation performance of SBX in the pyrite-catalyzed Fenton system was studied. The degradation conditions were as follows: [SBX] = 100 mg/L, [FeS2] = 0.5 g/L, and initial pH = 9.36 ± 0.5. As shown in Figure 2, the degradation efficiency of SBX increased with the increase in H2O2 concentration. When the H2O2 concentration was increased from 0.1 mM to 0.5 mM, the SBX degradation efficiency increased from 83.84% to 96.68% after 75 min of the reaction, which may be attributable to the increase in reactive groups (e.g., •OH) produced by the reaction of Fe2+ with H2O2 to promote the degradation of SBX. When the H2O2 concentration was further increased from 0.5 mM to 0.9 mM, the SBS removal rate did not change remarkably. Prolonging the reaction time favored the SBX degradation, and the degradation efficiency changed weakly when the reaction time exceeded 75 min. Thus, 0.5 mM H2O2 was selected as the optimal reaction concentration in the subsequent experiments.

3.1.2. The Effect of Pyrite Concentration

The effect of the pyrite concentration on the removal rate of SBX is displayed in Figure 3 (conditions: [SBX] = 100 mg/L, [H2O2] = 0.5 mM, and initial pH = 9.36 ± 0.5). As can be seen in Figure 3, the pyrite concentration played an important role in the removal rate of SBX. When the pyrite concentration increased from 0.1 g/L to 0.7 g/L, the removal rate of SBX increased during the reaction time ranging from approximately 5 to 45 min. With the further increase in the reaction time to 85 min, the removal rate of SBX with 0.7 g/L pyrite addition decreased to some extent. Notably, the removal rate of SBX with 0.9 g/L pyrite addition exhibited lower values than the corresponding results with 0.7 g/L pyrite addition. The results indicate that the excess addition of pyrite was detrimental to the SBX degradation. It is well accepted that the presence of pyrite can activate the decomposition of H2O2 [42]. This is rooted in the fact that pyrite provides iron ions and creates a heterogeneous Fenton system. The main chemical reactions involved are as follows [42,43]:
2FeS2 + 7O2 + 2H2O → 2Fe2+ + 4SO42− + 4H+
2FeS2 + 15H2O2 → 2Fe3+ + 4SO42−+ 2H+ +14H2O
FeS2 + 14Fe3+ + 8H2O → 15Fe2+ + 2SO42−+ 16H+
Fe2+ + H2O2→ Fe3+ + OH + •OH
Clearly, the presence of iron ions (Fe2+ and Fe3+) originating from pyrite oxidization by O2 and H2O2 enhanced the formation of •OH, which was responsible for the SBX degradation. However, an excess amount of Fe2+ in the Fenton system could react with the generated •OH through Equation (6), which would consume •OH and inhibit the oxidative degradation of SBX [44]. Thus, an appropriate concentration of pyrite (0.5 g/L) acted as a catalyst for the Fenton reaction and enhanced the xanthate degradation.
Fe2++ •OH → Fe3++ OH

3.1.3. The Effect of Initial SBX Concentration

The effect of initial SBX concentration on the degradation efficiency was further analyzed. As shown in Figure 4, it could be concluded that, generally, the SBX removal rate decreased with the increase in the initial SBX concentration at the same reaction time. It is possible that the oxidation of excess SBX molecules reached saturation and reduced the degradation capacity due to the limited reactive radicals generated by the Fenton reaction in the solution, whereas the initial SBX concentration had little effect on the final removal rate. For instance, even when the SBX concentration was increased to 150 mg/L, the removal rate reached 94.20% after a duration of 85 min. Therefore, considering the effective removal of SBX and reasonable practical application, the selected initial concentration of SBX was 100 mg/L.

3.2. Comparative Experiment Results

Two comparative experiments, (I) a H2O2-alone system and (II) a H2O2-FeS2 system, were conducted to further verify the specific role of pyrite in degrading SBX, and the results are shown in Figure 5. It can be seen that the SBX removal rate in the H2O2-FeS2 system reached 96.49% after 75 min under the recommended experimental conditions ([H2O2] = 0.5 mM, [FeS2] = 0.5 g/L, [SBX] = 100 mg/L, and initial pH = 9.36 ± 0.5), which was 12.85% higher than the corresponding result in the H2O2-alone system under the same conditions. The results provide strong support for the notion that the H2O2-alone system failed to oxidize SBX completely due to its limited oxidation capacity, while an appropriate concentration of pyrite could act as a catalyst to significantly improve the oxidation capacity of H2O2 by establishing a Fenton oxidation system. The utilization of pyrite in combination with H2O2 might be an efficient method for the degradation of SBX in flotation tailings wastewater.

3.3. SBX Degradation Mechanism in the Pyrite-Catalyzed Fenton System

3.3.1. The Identification of the Reactive Radical

EPR tests were conducted to analyze the generation of free radicals in the reaction system under the recommended experimental conditions. DMPO was used as a trapping agent of hydroxyl radicals because it reacted specifically with hydroxyl radicals and formed DMPO-•OH adducts [25]. In Figure 6a, no significant peaks appear in the spectra of the H2O2-alone system, indicating the low generating activity of hydroxyl radicals in this system. In the H2O2-FeS2 system, the adduct is observable as a strong EPR signal with four characteristic peak intensities of 1:2:2:1 for DMPO-•OH. The results confirm that the free radicals produced in the catalytic reaction were •OH. Thus, pyrite could effectively catalyze the decomposition of H2O2 and promoted •OH production.
Additionally, the role of reactive radicals (•OH) in the degradation SBX was analyzed based on the principle that TBA could capture hydroxyl radicals (•OH), and the results are shown in Figure 6b. The results confirm that the SBX in the H2O2-FeS2 system was almost completely degraded after 75 min of the reaction, while the SBX removal rate was about 72.38% after the addition of 200 mM TBA, which was 24.11% lower than the corresponding result without TBA addition. Notably, the SBX removal rate with TBA addition increased with the increase in the reaction time and presented relatively high levels (>68.49%) when the reaction time was more than 30 min. This phenomenon is mainly attributed to the adoption of SBX onto pyrite’s surfaces. This behavior only prompted the SBX transferring from the solution phase to the solid phase, and SBX failed to be mineralized to small molecules. However, this adoption inevitably caused electron transport and improved the reaction environment, which might have facilitated the interaction between SBX and H2O2. Thus, the reactive radicals (•OH) exerted an important role in the SBX degradation.

3.3.2. UV-Visible Spectrum Analysis Results

The UV full-scan spectrum of the SBX wastewater solution was obtained at a scan range of 200–400 nm and different reaction times (conditions: [H2O2] = 0.5 mM, [FeS2] = 0.5 g/L, [SBX] = 100 mg/L, and initial pH = 9.36 ± 0.5) with the aim of analyzing the SBX degradation process. As presented in Figure 7, the obvious absorption peaks located at 301 nm belong to the characteristic absorption peaks of SBX molecules [6]. The peak intensity of the SBX at 301 nm decreased gradually as the reaction time increased from 5 to 85 min. This result indicates that the residual concentration of SBX in the solution gradually decreased with the increasing reaction time. Furthermore, a new peak at 348 nm can be observed in the spectrum at different reaction times, which is assigned to butyl xanthate peroxide (BPX, C4H9OCS2O) [45]. The intensity of C4H9OCS2O presented a downtrend as the reaction time extended. This result indicates that SBX was degraded in the H2O2-FeS2 system, which was accompanied by the generation of an intermediate product (butyl xanthate peroxide). When the reaction exceeded 75 min, the peaks at 301 nm and 348 nm disappeared almost completely. Based on previous literature reports, it is known that SBX might be decomposed into SO42−, CO2, and H2O small-molecule substances in the degradation process [13,45].
However, the determination of SBX contaminants at 301 nm using UV-vis spectroscopy may have been interfered with by iron ions. To further analyze the degradation performance of SBX in the H2O2-FeS2 system, the COD of the solution samples was determined at different reaction times. As shown in Figure 7b, the COD removal rate increased with the increase in the reaction time within 75 min. When the reaction was carried out at 75 min, the COD removal rate reached 79%, which was possibly due to the further decomposition of the formed intermediates (i.e., C2H5OCS2O) into small inorganic molecules. When the reaction time was further extended, the COD removal rate did not change significantly, indicating that SBX and its intermediates were almost completely decomposed into small molecules such as SO42−, CO2, and H2O. This is in agreement with the results of the UV-visible analysis.

3.3.3. Solution pH and Iron Concentration Analysis

The solution pH and concentration variation in the iron ions in the pyrite-catalyzed Fenton system are shown in Figure 8. As shown in Figure 8a, the solution pH decreased gradually as the reaction time extended from 5 to 85 min. Specifically, the initial pH of 9.36 ± 0.5 decreased to a final pH of 7.4 ± 0.5. This was due to the fact that the SBX was decomposed into CO2, H2O, and SO42−. The concentration of total dissolved Fe exhibited an obvious increasing trend (Figure 8b). Meanwhile, the aqueous Fe2+ concentration remained at a low concentration without significant variation instead of running out, which ensured that the reaction proceeded continuously, indicating that aqueous Fe2+ was constantly released from the pyrite surface and that the Fe3+ recycle existed, as shown in Equation (5). Additionally, it could be concluded that the Fe3+ concentrations in the solution increased significantly because the aqueous Fe2+ generated from the pyrite surface was soon oxidized to Fe3+ used for the formation of •OH in the pyrite Fenton reaction. This result indicates that pyrite could significantly enhance the reactivity of a H2O2 system.
Along with the aforementioned results and discussion, we propose a reaction mechanism for SBX degradation via a pyrite-catalyzed Fenton system. As shown in Figure 9, the H2O2 adsorbed onto the pyrite surface and reacted with Fe2+, triggering the formation of reactive radicals (•OH) and Fe3+. The •OH attacked the SBX that adsorbed on the pyrite surface or existed in the solution and promoted the transformation of the SBX anion (C4H9OCS2) into the intermediate butyl xanthate peroxide (BPX, C4H9OCS2O). Finally, BPX intermediates were further oxidized to smaller products such as SO42− and the inorganic substances CO2 and H2O. Given the above results, the natural pyrite acted as a catalyst for the heterogeneous Fenton reaction and enhanced SBX degradation in the flotation tailings wastewater considerably.

4. Conclusions

The pyrite-catalyzed Fenton system exhibited excellent degradation efficiency for SBX, and under the same conditions, this system degraded SBX more efficiently than the H2O2 system. An appropriate concentration of pyrite acted as a catalyst to significantly improve the oxidation capacity of H2O2 by establishing a Fenton oxidation system. The concentration of total dissolved Fe exhibited an obvious increasing trend in the system and the aqueous Fe2+ concentration remained at a low concentration, indicating that aqueous Fe2+ was constantly released from the pyrite surface. Additionally, the results of EPR and quenching measurements suggested that hydroxyl radicals were the main active species in the H2O2-FeS2 system for the degradation of SBX. A possible reaction mechanism was proposed. H2O2 adsorbed onto the pyrite surface and reacted with Fe2+, triggering the formation of reactive radicals (•OH) and Fe3+. The •OH most likely attacked the SBX that adsorbed on the pyrite surface or existed in the solution and promoted the transformation of the SBX anion (C4H9OCS2) into the intermediate butyl xanthate peroxide (BPX, C4H9OCS2O). Finally, BPX intermediates were likely further oxidized to smaller products such as SO42−, CO2, and H2O under the ongoing attack of •OH.

Author Contributions

Data curation, writing—review and editing, X.G.; conceptualization, supervision, methodology, S.L.; investigation, methodology, J.Y.; investigation, methodology, Z.D.; methodology, formal analysis, A.Y.; conceptualization, supervision, S.W.; conceptualization, methodology, supervision, writing—review and editing, S.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 52164021); the Natural Science Foundation of Yunnan Province (Grant No. 2019FB078); and the Special Funds for this research project (Grant No. CCC21321119A and Grant No. CA22369M062A).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Xie, L.; Wang, J.; Lu, Q.; Hu, W.; Yang, D.; Qiao, C.; Peng, X.; Peng, Q.; Wang, T.; Sun, W.; et al. Surface interaction mechanisms in mineral flotation: Fundamentals, measurements, and perspectives. Adv. Colloid. Interface Sci. 2021, 295, 102491. [Google Scholar] [CrossRef] [PubMed]
  2. Cheng, H.; Lin, H.; Huo, H.; Dong, Y.; Xue, Q.; Cao, L. Continuous removal of ore floatation reagents by an anaerobic-aerobic biological filter. Bioresour. Technol. 2012, 114, 255–261. [Google Scholar] [CrossRef] [PubMed]
  3. Mhonde, N.; Johansson, L.-S.; Corin, K.; Schreithofer, N. The effect of sodium isobutyl xanthate on galena and chalcopyrite flotation in the presence of dithionite ions. Miner. Eng. 2021, 169, 106985. [Google Scholar] [CrossRef]
  4. Lou, J.; Lu, G.; Wei, Y.; Zhang, Y.; An, J.; Jia, M.; Li, M. Enhanced degradation of residual potassium ethyl xanthate in mineral separation wastewater by dielectric barrier discharge plasma and peroxymonosulfate. Sep. Purif. Technol. 2022, 282, 119955. [Google Scholar] [CrossRef]
  5. Amrollahi, A.; Massinaei, M.; Zeraatkar Moghaddam, A. Removal of the residual xanthate from flotation plant tailings using bentonite modified by magnetic nano-particles. Miner. Eng. 2019, 134, 142–155. [Google Scholar] [CrossRef]
  6. Zhou, P.; Shen, Y.; Zhao, S.; Chen, Y.; Gao, S.; Liu, W.; Wei, D. Hydrothermal synthesis of novel ternary hierarchical MoS2/TiO2/clinoptilolite nanocomposites with remarkably enhanced visible light response towards xanthates. Appl. Surf. Sci. 2021, 542, 148578. [Google Scholar] [CrossRef]
  7. Dong, J.; Liu, Q.; Yu, L.; Subhonqulov, S.H. The interaction mechanism of Fe3+ and NH4+ on chalcopyrite surface and its response to flotation separation of chalcopyrite from arsenopyrite. Sep. Purif. Technol. 2021, 256, 117778. [Google Scholar] [CrossRef]
  8. Yuan, J.; Li, S.; Ding, Z.; Li, J.; Yu, A.; Wen, S.; Bai, S. Treatment Technology and Research Progress of Residual Xanthate in Mineral Processing Wastewater. Minerals 2023, 13, 435–454. [Google Scholar] [CrossRef]
  9. Wang, L.; Fu, P.; Ma, Y.; Zhang, X.; Zhang, Y.; Yang, X. Steel slag as a cost-effective adsorbent for synergic removal of collectors, Cu(II) and Pb(II) ions from flotation wastewaters. Miner. Eng. 2022, 183, 107593. [Google Scholar] [CrossRef]
  10. Meng, X.; Wu, J.; Kang, J.; Gao, J.; Liu, R.; Gao, Y.; Wang, R.; Fan, R.; Khoso, S.A.; Sun, W.; et al. Comparison of the reduction of chemical oxygen demand in wastewater from mineral processing using the coagulation–flocculation, adsorption and Fenton processes. Miner. Eng. 2018, 128, 275–283. [Google Scholar] [CrossRef]
  11. Vučinić, D.R.; Lazić, P.M.; Rosić, A.A. Ethyl xanthate adsorption and adsorption kinetics on lead-modified galena and sphalerite under flotation conditions. Colloids Surf. A 2006, 279, 96–104. [Google Scholar] [CrossRef]
  12. Wei, C.; Qin, Z.; Wen-Qiang, M.A. A Research on Removal of Xanthate from Floatation Wastewater with Activated Carbon. Acta Petrol. Sin. 2010, 30, 262–267. [Google Scholar]
  13. Lin, H.; Qin, K.; Dong, Y.; Li, B. A newly-constructed bifunctional bacterial consortium for removing butyl xanthate and cadmium simultaneously from mineral processing wastewater: Experimental evaluation, degradation and biomineralization. J. Environ. Manag. 2022, 316, 115304. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, S.; Gong, W.; Mei, G.; Zhou, Q.; Bai, C.; Xu, N. Primary biodegradation of sulfide mineral flotation collectors. Miner. Eng. 2011, 24, 953–955. [Google Scholar] [CrossRef]
  15. Silvester, E.; Truccolo, D.; Hao, F.P. Kinetics and mechanism of the oxidation of ethyl xanthate and ethyl thiocarbonate by hydrogen peroxide. J. Am. Chem. Soc. 2002, 10, 1562–1571. [Google Scholar] [CrossRef]
  16. Lee, K.E.; Morad, N.; Teng, T.T.; Poh, B.T. Development, characterization and the application of hybrid materials in coagulation/flocculation of wastewater: A review. Chem. Eng. J. 2012, 203, 370–386. [Google Scholar] [CrossRef]
  17. Rezaei, R.; Massinaei, M.; Zeraatkar Moghaddam, A. Removal of the residual xanthate from flotation plant tailings using modified bentonite. Miner. Eng. 2018, 119, 1–10. [Google Scholar] [CrossRef]
  18. Arnold, S.M.; Hickey, W.J.; Harris, R.F. Degradation of Atrazine by Fenton’s Reagent: Condition Optimization and Product Quantification. Environ. Sci. Technol. 1995, 29, 2083–2089. [Google Scholar] [CrossRef]
  19. Neyens, E.; Baeyens, J. A review of classic Fenton’s peroxidation as an advanced oxidation technique. J. Hazard. Mater. 2003, 98, 33–50. [Google Scholar] [CrossRef]
  20. Vitale, P.; Ramos, P.B.; Colasurdo, V.; Fernandez, M.B.; Eyler, G.N. Treatment of real wastewater from the graphic industry using advanced oxidation technologies: Degradation models and feasibility analysis. J. Clean. Prod. 2019, 206, 1041–1050. [Google Scholar] [CrossRef]
  21. Tufail, A.; Price, W.E.; Mohseni, M.; Pramanik, B.K.; Hai, F.I. A critical review of advanced oxidation processes for emerging trace organic contaminant degradation: Mechanisms, factors, degradation products, and effluent toxicity. J. Water Process Eng. 2021, 40, 101778. [Google Scholar] [CrossRef]
  22. Duesterberg, C.; Waite, T. Process Optimization of Fenton Oxidation Using Kinetic Modeling. Environ. Sci. Technol. 2006, 40, 4189. [Google Scholar] [CrossRef] [PubMed]
  23. Kang, S.F.; Chang, H.M. Coagulation of textile secondary effluents with Fenton’s reagent. Water Sci. Technol. 1997, 36, 215–222. [Google Scholar] [CrossRef]
  24. Popovic, A.; Milicevic, S.; Milosevic, V.; Ivosevic, B.; Povrenovic, D. Fenton process in dispersed systems for industrial wastewater treatment. Hem. Ind. 2019, 73, 5. [Google Scholar] [CrossRef]
  25. Sun, Y.; Zhou, P.; Zhang, P.; Meng, S.; Zhou, C.; Liu, Y.; Zhang, H.; Xiong, Z.; Duan, X.; Lai, B. New insight into carbon materials enhanced Fenton oxidation: A strategy for green iron(III)/iron(II) cycles. Chem. Eng. J. 2022, 450, 138423. [Google Scholar] [CrossRef]
  26. Liang, C.; Bruell, C.J.; Marley, M.C.; Sperry, K.L. Persulfate oxidation for in situ remediation of TCE. I. Activated by ferrous ion with and without a persulfate-thiosulfate redox couple. Chemosphere 2004, 55, 1213–1223. [Google Scholar] [CrossRef] [PubMed]
  27. Pignatello, J.J.; Oliveros, E.; MacKay, A. Advanced Oxidation Processes for Organic Contaminant Destruction Based on the Fenton Reaction and Related Chemistry. Crit. Rev. Environ. Sci. Technol. 2006, 36, 1–84. [Google Scholar] [CrossRef]
  28. Teel, A.L.; Warberg, C.R.; Atkinson, D.A.; Watts, R.J. Comparison of mineral and soluble iron Fenton’s catalysts for the treatment of trichloroethylene. Water Res. 2001, 35, 977–984. [Google Scholar] [CrossRef]
  29. Arienzo, M. Oxidizing 2,4,6-trinitrotoluene with pyrite-H2O2 suspensions. Chemosphere 1999, 39, 1629–1638. [Google Scholar] [CrossRef]
  30. Katsoyiannis, I.A.; Ruettimann, T.; Hug, S.J. pH dependence of Fenton reagent generation and As(III) oxidation and removal by corrosion of zero valent iron in aerated water. Environ. Sci. Technol. 2008, 42, 7424–7430. [Google Scholar] [CrossRef]
  31. Matta, R.; Hanna, K.; Chiron, S. Fenton-like oxidation of 2,4,6-trinitrotoluene using different iron minerals. Sci. Total Environ. 2007, 385, 242–251. [Google Scholar] [CrossRef] [PubMed]
  32. Costa, R.C.C.; Moura, F.C.C.; Ardisson, J.D.; Fabris, J.D.; Lago, R.M. Highly active heterogeneous Fenton-like systems based on Fe0/Fe3O4 composites prepared by controlled reduction of iron oxides. Appl. Catal. B 2008, 83, 131–139. [Google Scholar] [CrossRef]
  33. Hu, X.B.; Deng, Y.H.; Gao, Z.Q.; Liu, B.Z.; Sun, C. Transformation and reduction of androgenic activity of 17α-methyltestosterone in Fe3O4/MWCNTs–H2O2 system. Appl. Catal. B 2012, 127, 167–174. [Google Scholar] [CrossRef]
  34. Moura, F.C.; Araujo, M.H.; Costa, R.C.; Fabris, J.D.; Ardisson, J.D.; Macedo, W.A.; Lago, R.M. Efficient use of Fe metal as an electron transfer agent in a heterogeneous Fenton system based on Fe0/Fe3O4 composites. Chemosphere 2005, 60, 1118–1123. [Google Scholar] [CrossRef] [PubMed]
  35. Diao, Z.; Liu, J.; Hu, Y.; Kong, L.; Jiang, D.; Xu, X. Comparative study of Rhodamine B degradation by the systems pyrite/H2O2 and pyrite/persulfate: Reactivity, stability, products and mechanism. Sep. Purif. Technol. 2017, 184, 374–383. [Google Scholar] [CrossRef]
  36. Rahimi, F.; van der Hoek, J.P.; Royer, S.; Javid, A.; Mashayekh-Salehi, A.; Jafari Sani, M. Pyrite nanoparticles derived from mine waste as efficient catalyst for the activation of persulfates for degradation of tetracycline. J. Water Process Eng. 2021, 40, 101808. [Google Scholar] [CrossRef]
  37. Zhang, Y.; Zhang, K.; Dai, C.; Zhou, X.; Si, H. An enhanced Fenton reaction catalyzed by natural heterogeneous pyrite for nitrobenzene degradation in an aqueous solution. Chem. Eng. J. 2014, 244, 438–445. [Google Scholar] [CrossRef]
  38. Che, H.; Bae, S.; Lee, W. Degradation of trichloroethylene by Fenton reaction in pyrite suspension. J. Hazard. Mater. 2011, 185, 1355–1361. [Google Scholar] [CrossRef]
  39. Lian, W.; Yi, X.; Huang, K.; Tang, T.; Wang, R.; Tao, X.; Zheng, Z.; Dang, Z.; Yin, H.; Lu, G. Degradation of tris(2-chloroethyl) phosphate (TCEP) in aqueous solution by using pyrite activating persulfate to produce radicals. Ecotoxicol. Environ. Saf. 2019, 174, 667–674. [Google Scholar] [CrossRef]
  40. Bae, S.; Lee, W. Enhanced reductive degradation of carbon tetrachloride by biogenic vivianite and Fe(II). Geochim. Cosmochim. Acta. 2012, 85, 170–186. [Google Scholar] [CrossRef]
  41. Bae, S.; Kim, D.; Lee, W. Degradation of diclofenac by pyrite catalyzed Fenton oxidation. Appl. Catal. B 2013, 134–135, 93–102. [Google Scholar] [CrossRef]
  42. Bonnissel-Gissinger, P.; Alnot, M.; Ehrhardt, J.J.; Behra, P. Surface Oxidation of Pyrite as a Function of pH. Environ. Sci. Technol. 1998, 32, 2839–2845. [Google Scholar] [CrossRef]
  43. Lipczynska-Kochany, E. Degradation of aqueous nitrophenols and nitrobenzene by means of the fenton reaction. Chemosphere 1991, 22, 529–536. [Google Scholar] [CrossRef]
  44. Chen, S.; Du, D. Degradation of n-butyl xanthate using fly ash as heterogeneous Fenton-like catalyst. J. Cent. S. Univ. 2014, 21, 1448–1452. [Google Scholar] [CrossRef]
  45. Li, G.; Teng, Q.; Sun, B.; Yang, Z.; Zhu, X. Synthesis scaly Ag-TiO2 loaded fly ash magnetic bead particles for treatment of xanthate wastewater. Colloids Surf. A 2021, 624, 126795. [Google Scholar] [CrossRef]
Figure 1. X-ray diffractograms of pyrite samples.
Figure 1. X-ray diffractograms of pyrite samples.
Minerals 13 01429 g001
Figure 2. Effect of H2O2 concentration on the removal rate of SBX (conditions: [SBX] = 100 mg/L, [FeS2] = 0.5 g/L, and initial pH = 9.36 ± 0.5).
Figure 2. Effect of H2O2 concentration on the removal rate of SBX (conditions: [SBX] = 100 mg/L, [FeS2] = 0.5 g/L, and initial pH = 9.36 ± 0.5).
Minerals 13 01429 g002
Figure 3. Effect of pyrite concentration on the removal rate of SBX (conditions: [SBX] = 100 mg/L, [H2O2] = 0.5 mM, and initial pH = 9.36 ± 0.5).
Figure 3. Effect of pyrite concentration on the removal rate of SBX (conditions: [SBX] = 100 mg/L, [H2O2] = 0.5 mM, and initial pH = 9.36 ± 0.5).
Minerals 13 01429 g003
Figure 4. Effect of initial SBX concentration on the removal rate of SBX (conditions: [H2O2] = 0.5 mM, [FeS2] = 0.5 g/L, and initial pH = 9.36 ± 0.5).
Figure 4. Effect of initial SBX concentration on the removal rate of SBX (conditions: [H2O2] = 0.5 mM, [FeS2] = 0.5 g/L, and initial pH = 9.36 ± 0.5).
Minerals 13 01429 g004
Figure 5. Effects of different oxidation treatments on SBX degradation (conditions: [H2O2] = 0.5 mM, [FeS2] = 0.5 g/L (if present), [SBX] = 100 mg/L, and initial pH = 9.36 ± 0.5).
Figure 5. Effects of different oxidation treatments on SBX degradation (conditions: [H2O2] = 0.5 mM, [FeS2] = 0.5 g/L (if present), [SBX] = 100 mg/L, and initial pH = 9.36 ± 0.5).
Minerals 13 01429 g005
Figure 6. Identification of reactive radical with EPR spectra in the H2O2-alone system and H2O2-FeS2 system during the SBX degradation process (a); the role of reactive radical (•OH) in the degradation SBX (b).
Figure 6. Identification of reactive radical with EPR spectra in the H2O2-alone system and H2O2-FeS2 system during the SBX degradation process (a); the role of reactive radical (•OH) in the degradation SBX (b).
Minerals 13 01429 g006
Figure 7. (a) UV-visible spectrum changes in SBX degradation in the pyrite-catalyzed Fenton system with time variations; (b) removal rate of COD at different reaction times (conditions: [H2O2] = 0.5 mM, [FeS2] = 0.5 g/L, [SBX] = 100 mg/L, and initial pH = 9.36 ± 0.5).
Figure 7. (a) UV-visible spectrum changes in SBX degradation in the pyrite-catalyzed Fenton system with time variations; (b) removal rate of COD at different reaction times (conditions: [H2O2] = 0.5 mM, [FeS2] = 0.5 g/L, [SBX] = 100 mg/L, and initial pH = 9.36 ± 0.5).
Minerals 13 01429 g007
Figure 8. Effects of reaction time on the solution pH (a) and on the concentrations of Fe2+ and total dissolved Fe (b) along with the degradation of SBX in the pyrite-catalyzed Fenton system (conditions: [H2O2] = 0.5 mM, [FeS2] = 0.5 g/L, [SBX] = 100 mg/L, and initial pH = 9.36 ± 0.5).
Figure 8. Effects of reaction time on the solution pH (a) and on the concentrations of Fe2+ and total dissolved Fe (b) along with the degradation of SBX in the pyrite-catalyzed Fenton system (conditions: [H2O2] = 0.5 mM, [FeS2] = 0.5 g/L, [SBX] = 100 mg/L, and initial pH = 9.36 ± 0.5).
Minerals 13 01429 g008
Figure 9. Proposed reaction mechanism for SBX degradation via pyrite-catalyzed Fenton system.
Figure 9. Proposed reaction mechanism for SBX degradation via pyrite-catalyzed Fenton system.
Minerals 13 01429 g009
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

Gong, X.; Li, S.; Yuan, J.; Ding, Z.; Yu, A.; Wen, S.; Bai, S. Natural Pyrite as a Catalyst for a Fenton Reaction to Enhance Xanthate Degradation in Flotation Tailings Wastewater. Minerals 2023, 13, 1429. https://doi.org/10.3390/min13111429

AMA Style

Gong X, Li S, Yuan J, Ding Z, Yu A, Wen S, Bai S. Natural Pyrite as a Catalyst for a Fenton Reaction to Enhance Xanthate Degradation in Flotation Tailings Wastewater. Minerals. 2023; 13(11):1429. https://doi.org/10.3390/min13111429

Chicago/Turabian Style

Gong, Xiang, Suqi Li, Jiaqiao Yuan, Zhan Ding, Anmei Yu, Shuming Wen, and Shaojun Bai. 2023. "Natural Pyrite as a Catalyst for a Fenton Reaction to Enhance Xanthate Degradation in Flotation Tailings Wastewater" Minerals 13, no. 11: 1429. https://doi.org/10.3390/min13111429

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