Next Article in Journal
Agri-Biodegradable Mulch Films Derived from Lignin in Empty Fruit Bunches
Next Article in Special Issue
Fast and Complete Destruction of the Anti-Cancer Drug Cytarabine from Water by Electrocatalytic Oxidation Using Electro-Fenton Process
Previous Article in Journal
Enhanced Performance and Stability of a Trimetallic CuZnY/SiBEA Catalyst in Ethanol to Butadiene Reaction by Introducing Copper to Optimize Acid/Base Ratio
Previous Article in Special Issue
Combining Ultraviolet Photolysis with In-Situ Electrochemical Oxidation for Degrading Sulfonamides in Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Accelerated Removal of Acid Orange 7 by Natural Iron Ore Activated Peroxymonosulfate System with Hydroxylamine for Promoting Fe(III)/Fe(II) Cycle

1
Faculty of Resources and Environmental Science, Hubei University, Wuhan 430062, China
2
Department of Environmental Science and Engineering, School of Resource and Environmental Sciences, Wuhan University, Wuhan 430079, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1148; https://doi.org/10.3390/catal12101148
Submission received: 15 August 2022 / Revised: 18 September 2022 / Accepted: 22 September 2022 / Published: 1 October 2022

Abstract

:
In this study, peroxymonosulfate (PMS) was activated by cheap and readily available natural iron ore to remove Acid Orange 7 (AO7) in water with the assistance of hydroxylamine (HA). Results show that the presence of HA could accelerate the Fe(II)/Fe(III) cycle on the ore surface, promoting the activation of PMS to generate reactive oxidative species. The effects of ore dosage, PMS dosage, HA dosage and initial pH on the degradation of AO7 were investigated in the HA/Ore/PMS system. Under the optimal conditions, the removal of AO7 could reach 93.1% during 30 min, which was 41.4% higher than the ore/PMS system. The AO7 removal increased with the increase of HA, PMS and ore dosage, but was unaffected by the initial solution pH. Based on radical scavenging experiments and EPR tests, the dominant reactive species in the HA/Ore/PMS system were revealed to be the sulfate radical (SO4•−), singlet oxygen (1O2), superoxide radical (O2•−) and hydroxyl radical (OH), which were responsible for the AO7 degradation. Furthermore, the possible reaction mechanism of PMS activation was proposed. This study provides an efficient technique for the removal of azo dye organic contaminant in water, which has great practical significance.

1. Introduction

Azo dyes are typical chemical products commonly used in the printing and dyeing industry [1]. Each year approximately 1.3 × 106 t of dyes are produced across the world, which inevitably results in some serious environmental pollution problems owing to the big discharge of dye wastewater into natural waters. The azo dyes wastewater with deep color, great capacity and high toxicity has already brought severe challenges to the existing water environment and human health. Additionally, this type of wastewater is bio-refractory and hard to be efficiently treated by conventional treatment technologies [2,3]. Therefore, it is necessary to develop an effective technology to efficiently remove the azo dye from the wastewater.
Sulfate-radical (SO4)-based advanced oxidation processes (AOPs) have been widely studied to remove refractory organic pollutants in wastewater. Of note, the oxidation potential (2.5–3.1 eV) of SO4•− is higher than that of the typical hydroxyl radical (OH, 1.9–2.7 eV) [4,5,6]. In addition, SO4•− has other advantages such as a longer half-life period and higher selectivity for pollutants decomposition, indicating a promising application prospect in water treatment. The solid oxidants peroxymonosulfate (PMS, HSO5) and peroxydisulfate (PDS, S2O82−) are the main sources of SO4•−. PMS and PDS are structurally similar to H2O2 and both of them have O-O bonds in their chemical structures. Nevertheless, the asymmetric structure of O-O bond in PMS makes it more active and easily decomposes into free radicals via the cleavage of O-O bonds during the PMS activation [7,8,9]. Based on this consideration, PMS activation processes have been widely investigated in wastewater treatment. There are many methods to activate PMS to produce reactive radicals for the removal of organics in water. For instance, homogeneous activation by transition metal ions (Co2+, Fe2+, Mn2+, Cu2+, etc.) is a commonly method for PMS activation [10]. However, this process requires an acidic pH environment and probably generates a large amount of sludge, causing the secondary pollution. In order to solve these problems, more and more studies are focusing on the heterogeneous activation process by using solid catalysts as activators [11,12,13,14,15]. Iron-based materials have been widely applied as active substances in Fenton-like reactions owing to their high activity, environmental friendliness and cost-effectiveness.
However, the synthesis process of iron-based compounds by chemicals is complex and costly. In recent decades, natural iron ores have come into view which are cheap and easy to obtain as peroxides activators. It is found that iron ore can activate persulfate to degrade organic contaminants, especially for groundwater and soil remediation, but the activation efficiency also needs to be improved. The main reason was attributed to the low efficiency of the active site Fe(II)/Fe(III) cycle [16]. It is considered that slow reduction of Fe(III) to Fe(II) would result in the accumulation of Fe(III) and the generation of Fe oxide sludge, finally inhibiting the formation of reactive oxygen species to degrade pollutants. To address this issue, a lot of studies proposed to apply some reductants to promote the reduction of Fe(III) [17,18,19,20,21]. The first reducing agent being applied was hydroxylamine (HA) that improved the pollutant removal in Fe(II)/persulfate and Fe(II)/H2O2 systems. For instance, Liu et al., employed HA to enhance the oxidation efficiency of sulfamethoxazole in the Fe(II)/PMS system. Compared with Fe(II)/PMS process, the HA/Fe(II)/PMS process showed about four times higher removal efficiency of sulfamethoxazole at pH 3.0 [21].
Further studies also demonstrated that HA displayed remarkable facilitation for the removal of pollutants in iron-based heterogeneous systems. Li reported that the Fe3O4/PMS system in the presence of HA could efficiently remove the herbicide atrazine under the near-neutral pH (5.0–6.8) (without buffer) [22], which was mainly due to the highly promoted Fe(III)/Fe(II) cycle by HA. It is noted that HA exists in the form of NH3OH+ under acidic conditions (pH < 5.96), and its reaction rate constant with reactive radicals (e.g., SO4•− and OH) is much lower than the rate constant of the reaction between the radicals and most of organic pollutants [23,24], thereby avoiding the senselessly consumption of reactive radicals by HA. In this regard, it is believed that the application of HA in ore activetedPMS system would also enhance the degradation of pollutants. To the best of our knowledge, a few related studies are reported but meaningful in consideration of the practical application by using nature iron ore.
In this work, the HA-enhanced Ore/PMS system was proposed to remove organic contamination in water. The acid orange 7 (AO7), a typical azo dye, was chosen as the target pollutant to study the performance of the HA/Ore/PMS process. The physical and chemical information of ore was characterized by a series of technologies including XRD, SEM, BET, XPS, and so on. The removal efficiencies of AO7 were investigated under the varied catalyst dosage, PMS dosage, HA dosage and initial pH, and the reaction kinetics were analyzed. Based on the radical quenching experiments and electron paramagnetic resonance (EPR) technique, the dominant reactive oxygen species that participated in the oxidation of AO7 have been further discussed and the possible mechanism of PMS activation over the natural ore was proposed. Briefly, This work provides a theoretical basis for the practical application of a natural ore activated PMS system in the remediation of water pollution.

2. Results and Discussion

2.1. Characterization of Ore

X-ray diffraction (XRD) analysis was conducted to investigate the solid phase composition of natural ore. As shown in Figure 1, three strong peaks can be observed at 27°, 35.5° and 45°, which are corresponding to the presence of Fe3O4, Fe3Si and SiO2, respectively. Some weak peaks were observed, indicating the possible existence of the mixture of iron, silicon and aluminum (FeAl2.7Si2.3). These results are in agreement with the component analysis in Section 3.1, which verified that the iron oxides are the main component of ore.
Morphologies of ore were also observed by SEM. The images at different magnifications are shown in Figure 2. It can be seen that the iron ore was irregularly lumpy with a rough surface. Most of the particles were tightly interconnected and the size distribution was not uniform. However, the surface elements seem to be fairly distributed as indicated in Figure 3. It is considered that the rough surface of ore might allow for more active sites for PMS activation and enhance its catalytic performance.
The specific surface area and pore composition of ore are important for the catalytic properties of the catalyst. As shown in Figure 4, the adsorption–desorption isotherm of the ore tends to be type IV (with H3 hysteresis loops) based on the International Union of Pure and Applied Chemistry (IUPAC) test guidelines, suggesting that the sample was a flake or fissure porous material [25]. When the P/P0 approached 1.0, there was a sharp increase in the adsorption isotherm, presumably due to the presence of large pores in the ore. However, the specific surface area of ore was low and determined to be 0.962 m2·s−1 with the average pore size being 6.805 nm (Table 1), which might be attributed to the tight aggregation of particles.

2.2. Effect of Different Reaction Systems on AO7 Degradation

To explore the oxidation capacity of the HA/Ore/PMS system, the removal of AO7 was comparatively carried out in different processes including HA/PMS, HA/Ore, Ore/PMS and HA/Ore/PMS (Figure 5a).
As can be seen in Figure 5a, there was almost no AO7 removal after 30 min in both HA/PMS (6.4%) and HA/Ore (1.8%) systems, indicating no adsorption of AO7 and negligible production of reactive oxygen species. The coupling of ore with PMS could remove AO7 by 51.7% under the experimental conditions, which was probably attributed to the generation of reactive radicals in the reaction between ore and PMS. In contrast, the removal of AO7 was greatly improved by introducing 0.05 mM HA in the Ore/PMS system and the removal efficiency of AO7 reached 93.1% within 30 min.It is assumed that the presence of Fe(II) on the ore surface could activate PMS to generate reactive radicals for degrading AO7 molecules. However, Fe(II) would gradually be oxidized to Fe(III) during the oxidation process, resulting in a lot of iron oxide sludge. Thus the degradation rate decreased with the reaction in the later stage. The removal of AO7 followed the pseudo-first-order kinetic (Figure 5b). The removal rate constant in HA/Ore/PMS process was found to be 0.091 min−1, about 4.33 times that in Ore/PMS system (0.021 min−1). As shown in Table 2, it can be seen that the natural ore is not inferior to these lab-made oxides in terms of degradation ability for organic contaminants, which demonstrates the high catalytic efficiency of the natural ore. The high reaction efficiency in the presence of HA could be ascribed to the significant enhancement of radical generation. The efficient Fe(III)/Fe(II) redox cycle is a vital factor in PMS activation because the rate constant value in Fe(III) reduction is much lower than that in Fe(II) oxidation. The presence of HA could realize the rapid reduction of Fe(III) to Fe(II), enhancing the PMS activation for reactive radicals generation and then improving the oxidation ability of the Ore/PMS system for the AO7 degradation. The total organic carbon (TOC) removal capacity of the HA/Ore/PMS system was also investigated and about 10% TOC removal was obtained after 90 min. The reason for the low TOC removal might be attributed to the small organic acids formed in the decomposition of AO7, which are more stubborn and difficult to be degraded.

2.3. Influencing Factors of HA/Ore/PMS System

2.3.1. Effect of Catalyst Dosage on AO7 Removal

The dosage of catalyst directly relates to the concentration of surface active sites Fe(II)/Fe(III) during the catalytic reaction, thus the effects of catalyst dosage (1 g/L, 2 g/L, 3 g/L) on the AO7 removal need to be investigated.
As shown in Figure 6, when the ore dosage increased from 1 g/L to 3 g/L, the removal efficiency increased with increasing catalyst dosage. The corresponding removal efficiencies of AO7 in 30 min were 81.3%, 93.1% and 96.0%, respectively. It is suggested that a larger amount of catalyst would bring more Fe(II)/Fe(III) to activate PMS. However, a further increase in ore dosage just caused a slight enhancement in the removal efficiency. The reasons might be due to the quenching of partial reactive radicals such as SO4•− and OH by excess Fe(II) through Equation (1), thereby reducing the oxidative efficiency of AO7 [28,29].
Fe(II) + SO4•−/OH → Fe(III) + SO42−/OH

2.3.2. Effects of PMS Dosage on AO7 Removal

Figure 7 shows the effect of PMS concentration on the removal of AO7 in the HA/Ore/PMS process. When the PMS dosage increased from 0.25 mM to 1 mM, the removal efficiency of AO7 increased from 46.8% to 90%, respectively. This suggested that the increased concentration of PMS would result in more PMS decomposition into reactive radicals for the AO7 degradation.
Nevertheless, further increase in PMS dosage did not improve the AO7 removal significantly, the removal of AO7 increased slowly to 93% at 2 mM PMS and 93.1% at 3 mM PMS, respectively. As PMS is the source of reactive species, an increase in PMS dosage can produce more reactive species to degrade contaminants. However it should be noted that excess PMS would in turn react with SO4•− and OH, finally decreasing the oxidation ability of the HA/Ore/PMS system.

2.3.3. Effects of HA Dosage on AO7 Removal

Figure 8 illustrates the effect of HA dosage on AO7 removal. When the dosage of HA was gradually increased from 0.025 mM to 0.1 mM, the AO7 removal efficiency increased accordingly. As previously reported, the HA as a strong reducing agent can effectively promote the cycle of Fe(II)/Fe(III) in the system and accelerate the generation of free radicals [22,30,31]. It should be worth noting that HA tends to be protonated to form NH3OH+ around pH 3.0, facilitating the Fe(III) reduction into Fe(II). Additionally, under acidic conditions, the reaction rate constants of HA with SO4•− and OH are much lower than that of the reaction between radicals and organic pollutants, as well as the reaction between Fe(II) and PMS in an aqueous solution. Consequently, an obvious enhanced removal of AO7 was observed.

2.3.4. Effect of Initial pH on AO7 Removal

It has been reported that the pH value is a crucial factor in the AOPs. The pH value not only affects the formation of active radicals but also affects the existence form of ion and substance species in solution [32]. Hence, it is necessary to evaluate the effect of solution pH on AO7 removal. Figure 9a displays the effect of different initial pH values on AO7 removal in the HA/Ore/HA system. The initial pH of the solution was adjusted to 3, 5, 7, and 9 with a diluted solution of sulfuric acid and sodium hydroxide.
As seen, the removal rates for AO7 were basically the same within 30 min under different initial pH conditions. In order to illustrate this phenomenon, the change of pH in solution during the oxidation process has been monitored, as shown in Figure 9b. It can be observed that all the pH rapidly dropped to around 3 after the addition of PMS. Thus, all the reactions with different initial pH are actually undergone in a similar pH environment, leading to almost the same removal efficiency of AO7.

2.4. Reaction Mechanism

In order to deeply study the mechanism of AO7 degradation, the experiments were conducted by using methanol (towards both SO4•− and OH), tert-butanol (towards OH), p-benzoquinone (towards superoxide radical, O2•−) and L-histidine (towards singlet oxygen, 1O2) as quenching reagents to explore the participated reactive oxygen species in the HA/Ore/PMS system [32,33].
As can be seen in Figure 10, the AO7 removal efficiency significantly dropped to 52.9%, 45.6%, 17.1% and 15.8% within 30 min in the presence of tert-butanol (400 mM), p-benzoquinone (10 mM), methanol (400 mM) and L-histidine (10 mM), respectively. This phenomenon may preliminarily suggest the simultaneous generation of SO4•− and 1O2 (primary), as well as OH and O2•− (secondary), participating in the degradation of AO7 in the HA/Ore/PMS system. Subsequently, EPR tests with DMPO as the spin trapping agent were carried out to further verify the generation of SO4•− and OH. As presented in Figure 11, distinct signals corresponding to DMPO-OH and DMPO-SO4 adducts (derived from OH and SO4•−) were observed in the Ore/PMS system, while their intensities were further enhanced with the presence of HA, providing clear evidence for the promoted generation of reactive radicals in HA/Ore/PMS system. Compared with the signal of DMPO-SO4, the peaks of the DMPO-OH adduct were more obvious, which might be due to the rapid transformation of DMPO-SO4 into DMPO-OH via nucleophilic substitution.
In the heterogeneous catalytic process, the generation of reactive species mainly occurs on the surface of catalyst. Thus, XPS spectra of the fresh and used ore are compared to understand the change of surface composition and chemical states before and after the reaction. Figure 12a shows Fe, O, Al, Si, Mg, Ca and C elements on the surface of the fresh and used samples, which are consistent with the results from EDS mapping. High-resolution XPS spectra of Fe 2p were displayed in Figure 12b, in which the surface Fe(II) and Fe(III) accounted for 46.07% and 53.93%, respectively, in the fresh catalyst based on the deconvolution of the Fe (2p) envelope. It is considered that once ore was used to activate PMS for the generation of reactive species, a large amount of Fe(II) would be quickly oxidized to Fe(III). However, the surface Fe(III) slightly increased from 53.93% to 54.64% on the used catalyst, probably indicating the enhanced reduction of Fe(III) by HA.
As shown in Figure 12c, three characteristic peaks located at 530.3 eV, 531.8 eV and 532.8 eV were attributed to the lattice oxygen (Fe-O), surface hydroxyl group (−OH) and the O in adsorbed H2O, respectively. For the fresh ore catalyst, a large number of the surface −OH was found on the surface, which may be beneficial for the catalytic oxidation process. After the oxidation of AO7, the percentage of the −OH component reduced greatly from 75.51% to 66.12%, probably because the surface −OH groups were also active sites in the activation of PMS.
Based on the above results and discussions, a possible reaction mechanism in the HA-enhanced Ore/PMS system was proposed in Figure 13. Remarkably, the ore surface readily adsorbs H2O molecules that tend to be dissociated into hydroxyl groups (−OH), which was consistent with the XPS results (Figure 12). For the PMS activation by iron ore, the H2O molecule was first physically adsorbed on the active sites of surface Fe(II) to form ≡Fe(II)−OH [22,32]. Then, ≡Fe(II)−OH reacts with PMS to form SO4•− and OH (Equations (2) and (3)). When the HA existed as NH3OH+ species in acidic conditions, it would react with the accumulated ≡Fe(III)-OH, promoting the reduction of ≡Fe(III)-OH to ≡Fe(II)-OH (Equations (4) and (5)), which in turn leads to an increase in PMS activation via Equation (2) to yield a large number of reactive oxidative species like SO4•−, 1O2, O2•− and OH (Equations (6)–(10)) [22,31,32,33,34,35]. Combined with the quenching experiments above, it is assumed that the SO4•− and 1O2 play a major role in the removal of AO7 in the HA/Ore/PMS system.
≡Fe(II)-OH + HSO5 → ≡Fe(III)-OH + SO4•− + OH
SO4•− + H2O → SO42− + OH + H+
≡Fe(III)-OH + NH3OH+ → ≡Fe(II)-OH + 1/2N2 + 2H+ + H2O
2≡Fe(III)-OH + NH3OH+ → 2≡Fe(II)-OH + 1/2N2O + 3H+ + 1/2H2O
≡Fe(III)-OH + HSO5 → SO5•− + H+ + ≡Fe(II)-OH
≡Fe(II)-OH + O2 → ≡Fe(III)-OH + O2•−
2SO5•− → 2SO42− + 1O2
2O2•− + 2H2O → H2O2 + 1O2 + 2OH
O2•− + OH → 1O2 + OH

2.5. Recyclability of Catalyst

The recycling tests were conducted to investigate the stability of iron ore in the oxidation process. The catalyst was easily separated, washed and dried at 100 °C for 12 h and finally maintained at room temperature for further use. As shown in Figure 14a, the removal efficiency was above 90.0% after 30 min of reaction in the 1st run. Though the AO7 removal efficiency decreased slightly after the first use, the removal efficiency was still relatively high (>65.0%) in the 2nd and 3rd runs. The decrease in efficiency might be due to the leaching of Fe ions, the adsorbed AO7 degradation intermediates on the catalyst surface and the conglomeration of the catalyst [31]. The leached Fe ions concentration in the solution has been monitored over multiple cycles (Figure 14b). Results indicated that the leaching Fe ions were slightly increased but remained below 0.2 mg/L in three cycles of experiments. All the values were lower than the legal limit (2.7 mg/L) imposed by the directives of the US Environmental Protection Agency, as well as the 10.0 mg/L of discharge limit (GB 8978-2002) regulated by the Chinese Environmental Protection Agency. Figure 14c shows the XRD of the used catalyst. It can be seen that the diffraction peaks at 27° and 45° (corresponding to SiO2 and Fe3O4) of used ore were much lower, indicating that Fe and Si may be released into the solution during the experiments, and the results are consistent with the iron leaching experiment.

3. Materials and Methods

3.1. Materials

The iron ore used in this experiment came from an iron mill in Huangshi City, Hubei Province. They were crushed into small particles by a vertical planetary ball mill (XQM-2, China), afterward poured into a mortar, ground for 10 min, and then passed through a 100 mesh sieve. Ore smaller than 100 mesh was taken for the experiment. The analysis results of the iron phase are shown in Table 3. The macroscale images of ore before and after pretreatment are shown in Figure 15.
Acid Orange 7 (AO7, AR) and potassium peroxymonosulfate (PMS, the active ingredient, KHSO5, content is 42–46%) were obtained from Macklin. Hydroxylamine hydrochloride (NH3OHCl, AR), methanol (CH4O, GR), tert-Butanol (C4H10O, AR), p-benzoquinone (C6H4O2, CP), L-Histidine (C6H9N3O2), 1,10-phenanthroline (C12H8N2, AR), sodium hydroxide (NaOH, AR) and sulfuric acid (H2SO4, AR) were purchased from Sinopharm Group. All of these chemicals were used as received without further purification and the solutions were prepared using deionized water.

3.2. Experimental Procedures

The AO7 degradation experiments were performed in 250 mL beakers at a temperature of 26 ± 0.5 °C with a constant stirring rate. Firstly, a stock solution of 200 mL AO7 (10 mg/L) was placed in a 250 mL glass beaker. Unless otherwise specified, the initial pH of the dye solution was adjusted by using 0.1 M sulfuric acid and 0.1 M sodium hydroxide. Subsequently, a certain amount of PMS and HA were added into the AO7 solution in turn and the required ore iron ore catalyst was added quickly to start timing. At given time intervals, 3.0 mL aliquots were filtered through the 0.45 μm membrane and immediately analyzed. All the experiments were carried out in duplicate.

3.3. Analytical Methods

The AO7 removal efficiency was calculated by measuring the absorbance at 485 nm (the maximum absorption wavelength of AO7) using a UV-9100 spectrophotometer. X-ray powder diffraction (XRD) on MiniFlex 600 (Rigaku, Japan) instrument with Cu Kα radiation. The morphology and composition of the samples were observed on a field emission scanning electron microscope (FESEM-EDS, SIGMA 500). N2 sorption isotherms were acquired from a physical adsorption apparatus (Quantachrome, (Boynton Beach, FL, USA)). The surface composition of the solid was characterized by X-ray photoelectron microscopy (XPS, Thermo Scientific 250Xi (Waltham, MA, USA)) under ultra-high vacuum (UHV) condition with Al-Kα X-ray. The total iron content analysis was carried out by 1,10-phenanthroline spectrophotometry, and the absorbance of solution containing iron ions was measured at 510 nm. EPR spectroscopy was used to identify the reactive species generated in the HA/Ore/PMS system with 5,5-dimethyl-1-pyrroline N-oxide (DMPO) as a radical-trapping agent. The concentration of TOC was measured by the vario TOC select (Elementar, Hanau, Germany).

4. Conclusions

In this study, HA was introduced as a reducing agent to accelerate Fe(III)/Fe(II) recycling on the iron ore surface, enhancing the iron ore activated PMS process for the degradation of AO7. In the HA/Ore/PMS system, the AO7 removal was improved with the increase of iron ore, PMS and HA dosages, but it needs to avoid the meaningless scavenging and wasting by the excessive dose. Under the optimal conditions (Ore = 2 g/L, PMS = 2 mM, HA = 0.05 mM), the removal efficiency of AO7 (10 mg/L) could reach 93.1% within 30 min, which demonstrates the high efficiency of the natural ore in PMS activation for organics degradation. Furthermore, it is verified that the reactive species including SO4•−, 1O2, O2•− and OH participated in the oxidation process, in which SO4•− and 1O2 play a major role in AO7 removal. In conclusion, the HA/Ore/PMS system is an efficient method for the remediation of azo dye-containing wastewater.

Author Contributions

Conceptualization, Y.X.; methodology, D.H.; software, J.C. and W.H.; validation, D.H.; formal analysis, D.H. and L.W.; investigation, D.H.; resources, H.C. and H.Z.; data curation, D.H.; writing—original draft preparation, D.H. and L.W.; writing—review and editing, F.L. and Y.X.; visualization, D.H. and L.W.; supervision, H.L. and Y.X.; project administration, Y.X.; funding acquisition, H.L. 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. 22006034) and the Hubei Provincial Natural Science Foundation of China (Grant No. 2020CFB485).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author (Y.X.), upon reasonable request.

Acknowledgments

Authors gratefully acknowledge the financial support from the National Natural Science Foundation of China and the Hubei Provincial Natural Science Foundation of China.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chung, K.-T. Azo dyes and human health: A review. J. Environ. Sci. Health Part C 2016, 34, 233–261. [Google Scholar] [CrossRef] [PubMed]
  2. Solís, M.; Solís, A.; Pérez, H.I.; Manjarrez, N.; Flores, M. Microbial decolouration of azo dyes: A review. Process Biochem. 2012, 47, 1723–1748. [Google Scholar] [CrossRef]
  3. Faryal, R.; Hameed, A. Isolation and Characterization of various fungal strains from textile effluent for their use in bioremediation. Pak. J. Bot. 2005, 37, 1003–1008. [Google Scholar]
  4. Yang, Q.; Choi, H.; Chen, Y.; Dionysiou, D.D. Heterogeneous Activation of peroxymonosulfate by supported cobalt catalysts for the degradation of 2,4-Dichlorophenol in water: The Effect of support, cobalt precursor, and UV Radiation. Appl. Catal. B Environ. 2008, 77, 300–307. [Google Scholar] [CrossRef]
  5. Mahdi Ahmed, M.; Barbati, S.; Doumenq, P.; Chiron, S. Sulfate radical anion oxidation of diclofenac and sulfamethoxazole for water decontamination. Chem. Eng. J. 2012, 197, 440–447. [Google Scholar] [CrossRef]
  6. Guan, Y.-H.; Ma, J.; Li, X.-C.; Fang, J.-Y.; Chen, L.-W. Influence of PH on the formation of sulfate and hydroxyl radicals in the UV/Peroxymonosulfate system. Environ. Sci. Technol. 2011, 45, 9308–9314. [Google Scholar] [CrossRef]
  7. Zhou, Z.; Liu, X.; Sun, K.; Lin, C.; Ma, J.; He, M.; Ouyang, W. Persulfate-based advanced oxidation processes (AOPs) for organic-contaminated soil remediation: A review. Chem. Eng. J. 2019, 372, 836–851. [Google Scholar] [CrossRef]
  8. Li, Y.; Liu, L.-D.; Liu, L.; Liu, Y.; Zhang, H.-W.; Han, X. Efficient oxidation of phenol by persulfate using manganite as a catalyst. J. Mol. Catal. Chem. 2016, 411, 264–271. [Google Scholar] [CrossRef]
  9. Betterton, E.A.; Hoffmann, M.R. Kinetics and mechanism of the oxidation of aqueous hydrogen sulfide by peroxymonosulfate. Environ. Sci. Technol. 1990, 24, 1819–1824. [Google Scholar] [CrossRef]
  10. Anipsitakis, G.P.; Dionysiou, D.D. Radical generation by the interaction of transition metals with common oxidants. Environ. Sci. Technol. 2004, 38, 3705–3712. [Google Scholar] [CrossRef]
  11. Zhu, C.; Fang, G.; Dionysiou, D.D.; Liu, C.; Gao, J.; Qin, W.; Zhou, D. Efficient transformation of DDTs with persulfate activation by zero-valent iron nanoparticles: A mechanistic study. J. Hazard. Mater. 2016, 316, 232–241. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, S.; Xiong, P.; Zhan, W.; Xiong, L. Degradation of ethylthionocarbamate by pyrite-activated persulfate. Miner. Eng. 2018, 122, 38–43. [Google Scholar] [CrossRef]
  13. Yu, B.; Li, Z.; Zhang, S. Zero-valent copper-mediated peroxymonosulfate activation for efficient degradation of azo dye orange G. Catalysts 2022, 12, 700. [Google Scholar] [CrossRef]
  14. Salama, R.S.; El-Bahy, S.M.; Mannaa, M.A. Sulfamic acid supported on mesoporous MCM-41 as a Novel, efficient and reusable heterogenous solid acid catalyst for synthesis of xanthene, dihydropyrimidinone and coumarin derivatives. Colloids Surf. Physicochem. Eng. Asp. 2021, 628, 127261. [Google Scholar] [CrossRef]
  15. Altass, H.M.; Morad, M.; Khder, A.E.-R.S.; Mannaa, M.A.; Jassas, R.S.; Alsimaree, A.A.; Ahmed, S.A.; Salama, R.S. Enhanced catalytic activity for CO oxidation by highly active Pd nanoparticles supported on reduced graphene oxide /copper metal organic framework. J. Taiwan Inst. Chem. Eng. 2021, 128, 194–208. [Google Scholar] [CrossRef]
  16. Wang, Y.R.; Chu, W. Degradation of a xanthene dye by Fe(II)-mediated activation of oxone process. J. Hazard. Mater. 2011, 186, 1455–1461. [Google Scholar] [CrossRef]
  17. Bengtsson, G.; Fronaeus, S.; Bengtsson-Kloo, L. The kinetics and mechanism of oxidation of hydroxylamine by iron(III). J. Chem. Soc. Dalton Trans. 2002, 12, 2548–2552. [Google Scholar] [CrossRef]
  18. Yin, R.; Hu, L.; Xia, D.; Yang, J.; He, C.; Liao, Y.; Zhang, Q.; He, J. Hydroxylamine promoted Fe(III)/Fe(II) cycle on ilmenite surface to enhance persulfate catalytic activation and aqueous pharmaceutical ibuprofen degradation. Catal. Today 2020, 358, 294–302. [Google Scholar] [CrossRef]
  19. Johnson, M.D.; Hornstein, B.J. The kinetics and mechanism of the ferrate(VI) oxidation of hydroxylamines. Inorg. Chem. 2003, 42, 6923–6928. [Google Scholar] [CrossRef]
  20. Wang, J.; Zhang, M.; Zhou, R.; Li, J.; Zhao, W.; Chen, W. Trace Cu(II) can enhance the degradation of orange II in Fe(II)/hydroxylamine/persulfate system. J. Environ. Chem. Eng. 2021, 9, 104907. [Google Scholar] [CrossRef]
  21. Liu, G.; Li, X.; Han, B.; Chen, L.; Zhu, L.; Campos, L.C. Efficient degradation of sulfamethoxazole by the Fe(II)/HSO5- process enhanced by hydroxylamine: Efficiency and mechanism. J. Hazard. Mater. 2017, 322, 461–468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Li, J.; Wan, Y.; Li, Y.; Yao, G.; Lai, B. Surface Fe(III)/Fe(II) cycle promoted the degradation of atrazine by peroxymonosulfate activation in the presence of hydroxylamine. Appl. Catal. B Environ. 2019, 256, 117782. [Google Scholar] [CrossRef]
  23. Neta, P.; Huie, R.E.; Ross, A.B. Rate constants for reactions of inorganic radicals in aqueous solution. J. Phys. Chem. Ref. Data 1988, 17, 1027–1284. [Google Scholar] [CrossRef]
  24. Buxton, G.V.; Greenstock, C.L.; Phillips Helman, W.; Ross, A.B.; Tsang, W. Critical review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals (ṠOH/ṠO- in aqueous solution. J. Phys. Chem. Ref. Data 1988, 17, 513–886. [Google Scholar] [CrossRef] [Green Version]
  25. Naderi, M. Chapter fourteen—Surface area: Brunauer–emmett–teller (BET). In Progress in Filtration and Separation; Tarleton, S., Ed.; Academic Press: Oxford, UK, 2015; pp. 585–608. ISBN 978-0-12-384746-1. [Google Scholar]
  26. Sang, W.; Li, Z.; Huang, M.; Wu, X.; Li, D.; Mei, L.; Cui, J. Enhanced transition metal oxide based peroxymonosulfate activation by hydroxylamine for the degradation of sulfamethoxazole. Chem. Eng. J. 2020, 383, 123057. [Google Scholar] [CrossRef]
  27. Oh, D.; Lee, C.-S.; Kang, Y.-G.; Chang, Y.-S. Hydroxylamine-assisted Peroxymonosulfate activation using cobalt ferrite for sulfamethoxazole degradation. Chem. Eng. J. 2020, 386, 123751. [Google Scholar] [CrossRef]
  28. Kusic, H.; Peternel, I.; Ukic, S.; Koprivanac, N.; Bolanca, T.; Papic, S.; Bozic, A.L. Modeling of Iron activated persulfate oxidation treating reactive azo dye in water matrix. Chem. Eng. J. 2011, 172, 109–121. [Google Scholar] [CrossRef]
  29. Bu, L.; Shi, Z.; Zhou, S. Modeling of Fe(II)-activated persulfate oxidation using atrazine as a target contaminant. Sep. Purif. Technol. 2016, 169, 59–65. [Google Scholar] [CrossRef]
  30. Hughes, M.N.; Nicklin, H.G.; Shrimanker, K. Autoxidation of hydroxylamine in alkaline solutions. Part II. kinetics. the acid dissociation constant of hydroxylamine. J. Chem. Soc. Inorg. Phys. Theor. 1971, 3485–3487. [Google Scholar] [CrossRef]
  31. Li, Z.-Y.; Wang, L.; Liu, Y.-L.; Zhao, Q.; Ma, J. Unraveling the interaction of hydroxylamine and Fe(III) in Fe(II)/persulfate system: A kinetic and simulating study. Water Res. 2020, 168, 115093. [Google Scholar] [CrossRef]
  32. Xu, Y.; Ai, J.; Zhang, H. The Mechanism of degradation of bisphenol a using the magnetically separable CuFe2O4/peroxymonosulfate heterogeneous oxidation process. J. Hazard. Mater. 2016, 309, 87–96. [Google Scholar] [CrossRef] [PubMed]
  33. Yu, H.; Liu, Y.; Xu, M.; Cong, S.; Liu, M.; Zou, D. Hydroxylamine facilitated heterogeneous fenton-like reaction by nano micro-electrolysis material for rhodamine B degradation. J. Clean. Prod. 2021, 316, 128136. [Google Scholar] [CrossRef]
  34. Lou, X.; Fang, C.; Geng, Z.; Jin, Y.; Xiao, D.; Wang, Z.; Liu, J.; Guo, Y. Significantly enhanced base activation of peroxymonosulfate by polyphosphates: Kinetics and mechanism. Chemosphere 2017, 173, 529–534. [Google Scholar] [CrossRef] [PubMed]
  35. Khan, A.U.; Kasha, M. Singlet molecular oxygen in the haber-weiss reaction. Proc. Natl. Acad. Sci. USA 1994, 91, 12365–12367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. XRD spectra of the ore sample.
Figure 1. XRD spectra of the ore sample.
Catalysts 12 01148 g001
Figure 2. SEM images of raw ore at different magnifications. (a) Mag = 297X. (b) Mag = 2.00KX. (c) Mag = 10.00KX.
Figure 2. SEM images of raw ore at different magnifications. (a) Mag = 297X. (b) Mag = 2.00KX. (c) Mag = 10.00KX.
Catalysts 12 01148 g002
Figure 3. EDS images of raw ore (Pt element comes from the gold sprayed procedure).
Figure 3. EDS images of raw ore (Pt element comes from the gold sprayed procedure).
Catalysts 12 01148 g003
Figure 4. N2 adsorption−desorption isotherms and pore size distribution of iron ore.
Figure 4. N2 adsorption−desorption isotherms and pore size distribution of iron ore.
Catalysts 12 01148 g004
Figure 5. (a) AO7 removal in different systems, (b) kinetic analysis for the AO7 removal in different systems ([AO7] = 10 mg/L, [Ore] = 2 g/L, [PMS] = 2 mM, [HA] = 0.05 mM, initial pH = 7).
Figure 5. (a) AO7 removal in different systems, (b) kinetic analysis for the AO7 removal in different systems ([AO7] = 10 mg/L, [Ore] = 2 g/L, [PMS] = 2 mM, [HA] = 0.05 mM, initial pH = 7).
Catalysts 12 01148 g005
Figure 6. Effect of catalyst dosage on the AO7 removal ([AO7] = 10 mg/L, [PMS] = 2 mM, [HA] = 0.05 mM, initial pH = 7).
Figure 6. Effect of catalyst dosage on the AO7 removal ([AO7] = 10 mg/L, [PMS] = 2 mM, [HA] = 0.05 mM, initial pH = 7).
Catalysts 12 01148 g006
Figure 7. Effect of PMS dosage on the AO7 removal ([AO7] = 10 mg/L, [Ore] = 2 g/L, [HA] = 0.05 mM, initial pH = 7).
Figure 7. Effect of PMS dosage on the AO7 removal ([AO7] = 10 mg/L, [Ore] = 2 g/L, [HA] = 0.05 mM, initial pH = 7).
Catalysts 12 01148 g007
Figure 8. Effect of HA dosage on the AO7 removal ([AO7] = 10 mg/L, [Ore] = 2 g/L, [PMS] = 2 mM, initial pH = 7).
Figure 8. Effect of HA dosage on the AO7 removal ([AO7] = 10 mg/L, [Ore] = 2 g/L, [PMS] = 2 mM, initial pH = 7).
Catalysts 12 01148 g008
Figure 9. (a) Effect of initial solution pH on AO7 removal. (b) pH changes in different systems ([AO7] = 10 mg/L, [Ore] = 2 g/L, [PMS] = 2 mM, [HA] =0.05 mM).
Figure 9. (a) Effect of initial solution pH on AO7 removal. (b) pH changes in different systems ([AO7] = 10 mg/L, [Ore] = 2 g/L, [PMS] = 2 mM, [HA] =0.05 mM).
Catalysts 12 01148 g009
Figure 10. Effects of quenching reagents on AO7 removal in HA/Ore/PMS system ([AO7] = 10 mg/L, [Ore] = 2 g/L, [PMS] = 2 mM, [HA] = 0.05 mM, initial pH = 7).
Figure 10. Effects of quenching reagents on AO7 removal in HA/Ore/PMS system ([AO7] = 10 mg/L, [Ore] = 2 g/L, [PMS] = 2 mM, [HA] = 0.05 mM, initial pH = 7).
Catalysts 12 01148 g010
Figure 11. EPR spectra of DMPO−OH and DMPO-SO4 in the Ore/PMS and HA/Ore/PMS systems.
Figure 11. EPR spectra of DMPO−OH and DMPO-SO4 in the Ore/PMS and HA/Ore/PMS systems.
Catalysts 12 01148 g011
Figure 12. (a) Wide survey XPS spectra, (b) Fe 2p XPS and (c) O 1s XPS envelop of the fresh and used ore catalyst.
Figure 12. (a) Wide survey XPS spectra, (b) Fe 2p XPS and (c) O 1s XPS envelop of the fresh and used ore catalyst.
Catalysts 12 01148 g012
Figure 13. The reaction mechanism of the HA/Ore/PMS system.
Figure 13. The reaction mechanism of the HA/Ore/PMS system.
Catalysts 12 01148 g013
Figure 14. (a) Catalyst recycling experiments ([AO7] = 10 mg/L, [Ore] = 2 g/L, [PMS] = 2 mM, [HA] = 0.05 mM, initial pH = 7), (b) leaching of Fe ions over the multiple cycles. (c) XRD spectra of the fresh and used ore sample.
Figure 14. (a) Catalyst recycling experiments ([AO7] = 10 mg/L, [Ore] = 2 g/L, [PMS] = 2 mM, [HA] = 0.05 mM, initial pH = 7), (b) leaching of Fe ions over the multiple cycles. (c) XRD spectra of the fresh and used ore sample.
Catalysts 12 01148 g014
Figure 15. (a) Natural raw ore before pretreatment, (b) natural raw ore after pretreatment.
Figure 15. (a) Natural raw ore before pretreatment, (b) natural raw ore after pretreatment.
Catalysts 12 01148 g015
Table 1. The specific surface areas results for the iron ore.
Table 1. The specific surface areas results for the iron ore.
Iron OreValues
Specific surface area (BET)/(m2 g−1)0.962
Average pore volume/(cm3 m−1)0.016
Average pore size/(nm)6.805
Table 2. The removal of various pollutants by different catalysts in the HA-enhanced system.
Table 2. The removal of various pollutants by different catalysts in the HA-enhanced system.
Target PollutantActivation SystemRemoval Rate (%)k (min−1)References
AtrazineHA/Fe3O4/PMS940.152[22]
SulfamethoxazoleHA/Fe2O3/PMS90.80.077[26]
SulfamethoxazoleHA/CoFe2O4/PMS1000.034[27]
AO7HA/Ore/PMS93.10.091This work
Table 3. The iron phase analysis results of a comprehensive sample of raw ore.
Table 3. The iron phase analysis results of a comprehensive sample of raw ore.
Phase NameFe3O4Fe2(CO3)3Fe2O3FeSFe2(SiO3)3Full Iron
Content (%)24.470.873.151.740.3530.58
Distribution rate (%)80.022.8510.295.701.14100.00
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, H.; Wang, L.; He, D.; Cai, J.; He, W.; Liu, F.; Chen, H.; Zhang, H.; Xu, Y. Accelerated Removal of Acid Orange 7 by Natural Iron Ore Activated Peroxymonosulfate System with Hydroxylamine for Promoting Fe(III)/Fe(II) Cycle. Catalysts 2022, 12, 1148. https://doi.org/10.3390/catal12101148

AMA Style

Li H, Wang L, He D, Cai J, He W, Liu F, Chen H, Zhang H, Xu Y. Accelerated Removal of Acid Orange 7 by Natural Iron Ore Activated Peroxymonosulfate System with Hydroxylamine for Promoting Fe(III)/Fe(II) Cycle. Catalysts. 2022; 12(10):1148. https://doi.org/10.3390/catal12101148

Chicago/Turabian Style

Li, Haibo, Linfeng Wang, Dingyuan He, Jie Cai, Wenjie He, Fuzhen Liu, Hanxiao Chen, Hui Zhang, and Yin Xu. 2022. "Accelerated Removal of Acid Orange 7 by Natural Iron Ore Activated Peroxymonosulfate System with Hydroxylamine for Promoting Fe(III)/Fe(II) Cycle" Catalysts 12, no. 10: 1148. https://doi.org/10.3390/catal12101148

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