Next Article in Journal
Influence of the LaMn1−xFexO3 (x = 0–1) Composition on Catalytic Activities in the Reactions Involving Oxygen
Next Article in Special Issue
Design of BiOCl/WO3@Polyaniline Organic–Inorganic Nanocomposite Photocatalyst for the Efficient Decontamination of 2-Chlorophenol from Wastewater
Previous Article in Journal
Promoting the Electrocatalytic Ethanol Oxidation Activity of Pt by Alloying with Cu
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Wet Peroxide Oxidation of Anionic Pollutants over Fluorinated Fe3O4 Microspheres at Circumneutral pH Values

1
State Key Laboratory of New Textile Materials & Advanced Processing Technologies, Wuhan Textile University, Wuhan 430200, China
2
School of Chemistry and Environmental Engineering, Wuhan Institute of Technology, Wuhan 430205, China
3
Henan Institute of Advanced Technology, Zhengzhou University, Zhengzhou 450002, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(12), 1564; https://doi.org/10.3390/catal12121564
Submission received: 7 November 2022 / Revised: 22 November 2022 / Accepted: 30 November 2022 / Published: 2 December 2022
(This article belongs to the Special Issue Nanocatalysts for the Degradation of Refractory Pollutants)

Abstract

:
Fluorinated Fe3O4 microspheres with 7.1 ± 1.4 wt% of fluoride (F-Fe3O4-1) were prepared via glycothermal synthesis. Fluorination significantly enhanced the activity of F-Fe3O4-1 in catalytic wet peroxide oxidation of anionic dyes (including orange G (OG) and congo red) at pH ~7. However, the promotional effect of fluorination became less obvious for amphoteric rhodamine B and was not observed for cationic methylene blue. After reacting with H2O2 (40 mM) for 2 h at pH 6.5 and 40 °C, the decolorization rates of OG (0.1 mM) and the pseudo-first-order rate constant were 96.8% and 0.0284 min−1 over F-Fe3O4-1 versus 17.6% and 0.0011 min−1 over unmodified Fe3O4. The effects of reaction parameters (initial H2O2 concentration and pH value and reaction temperature) on OG decolorization with H2O2 over F-Fe3O4-1 were investigated. The reusability of F-Fe3O4-1 was demonstrated by OG decolorization in eight consecutive runs. Fluorination increased the isoelectric point of F-Fe3O4-1 to 8.7 and facilitated the adsorption and degradation of anionic dyes on the surface of F-Fe3O4-1 at pH ~7. Scavenging tests and EPR spectra supported that hydroxyl radicals were the main reactive species for the OG decolorization over F-Fe3O4-1.

Graphical Abstract

1. Introduction

Organic compounds constitute one of the most common pollutants found in wastewater that pose potential hazards to the ecosystem and the environment. Fenton oxidation is effective for organic wastewater remediation since it can generate strongly oxidative hydroxyl radicals (•OH, E0 = 2.80 V) that are capable of degrading most recalcitrant organic pollutants. However, the Fenton process requires acidic condition (pH ~3) and generates undesired iron sludge. Catalytic wet peroxide oxidation (CWPO) may overcome these problems since it can work at circumneutral pH values and avoid the formation of iron sludge [1,2,3,4]. However, most heterogeneous Fenton-like catalysts exhibit lower activity than the conventional homogeneous Fenton process [2,5,6,7,8] since the kinetics of a heterogeneous catalytic reaction are also affected by the mass transfer process inside the reaction system (e.g., adsorption and diffusion) besides the intrinsic activity of a catalyst. In this regard, the development of highly active and stable heterogeneous Fenton-like catalysts with favorable mass transfer is critical for a successful CWPO process.
Magnetite (Fe3O4) is a promising heterogeneous Fenton-like catalyst owing to its intrinsic peroxidase-like activity and efficient magnetic separation [2,9,10,11,12]. However, it exhibits very low catalytic activity for the degradation of anionic organic contaminants under circumneutral pH conditions [10,13]. One reason for this is its low isoelectric point (IEP, about 6–7) [14,15,16]. As such, Fe3O4 particles are negatively charged at pH ~7 and cannot effectively adsorb anionic contaminants due to electrostatic repulsion [14]. In addition, the activity of Fe3O4 particles in activating H2O2 under mild conditions (e.g., pH ~7 and room temperature) is still too weak and has to be further improved for practical applications [2].
Doping is a simple and useful methodology to enhance the activity of magnetite, wherein metallic dopants (e.g., Ti4+, Cu2+, and Cr3+) are the most commonly used [2,6,14,17,18]. In sharp contrast, nonmetal-doped Fe3O4 have been seldom investigated. Fluoride doping is known for modulating semiconductor band structures with improved photocatalytic performances (e.g., F-doped TiO2, SnO2, Co3O4, γ-Fe2O3, Cu2O, and Bi2WO6 [19,20,21,22,23,24]). In fluorinated Fe3O4 (F-Fe3O4) microspheres [25], fluoride may replace lattice oxygen or coordinate with surface lattice iron via ligand exchange with hydroxide groups. The substitution of lattice O2− with F results in a framework with extra positive charge due to a charge imbalance. A number of possibilities exist for charge compensation, including by attracting anions (e.g., OH, F, and Cl) from the solution, by reducing the average valence of the iron cations (e.g., increasing the content of low-valence ferrous (Fe2+) ions), or by introducing some cationic vacancies. Therefore, fluorination presents largely unexplored opportunities for adjusting the physicochemical properties of F-Fe3O4 for specific applications, e.g., it is expected to increase the surface zeta potential of the pristine Fe3O4 for the enhanced adsorption of anionic compounds via electrostatic attraction and to improve the H2O2-activation capability via the Fenton-like reaction (Equation (1)).
≡Fe2+ + H2O2 → ≡Fe3+ + ·OH + OH (≡ represents the particle surface)
In this work, we found that in comparison with unmodified Fe3O4, fluorination significantly enhanced the catalytic activity of F-Fe3O4 microspheres for the degradation of anionic pollutants (e.g., orange G (OG) and congo red (CR) in Scheme 1). However, this promotional effect became less obvious for an amphoteric dye rhodamine B (RhB), and no promotion was observed for a cationic dye methylene blue (MB)). The different activities of F-Fe3O4 in the degradation of cationic and anionic dyes were mainly attributed to the adsorption-enhanced catalytic oxidation of anionic dyes on its surface via a free-radical reaction pathway.

2. Results and Discussion

2.1. Catalytic Performances of F-Fe3O4-r Microspheres

As shown in Figure 1a, fluorination significantly enhanced the decolorization rate (DR) of OG (0.1 mM) with H2O2 (40 mM) after reacting for 2 h at pH 6.5 and 40 °C over F-Fe3O4-r (e.g., the 2 h DR increased from 17.6% for Fe3O4-blank to 62.2% for F-Fe3O4-0.1). The best performance (96.8% of DR after 2 h) was obtained over F-Fe3O4-1. A further increase in the fluoride content resulted in a gradual decrease in the 2 h DRs of OG (0.1 mM) (68.4% for F-Fe3O4-2 and 47.5% for F-Fe3O4-3), which may be related to the increase in the Na3FeF6 impurity in F-Fe3O4-r (r = 2, 3) [25]. By assuming that the instantaneous concentration of •OH is constant since the amount of H2O2 is in larger excess with regard to OG (40 mM vs. 0.1 mM), the decolorization process of OG in water was well fitted with the pseudo-first-order kinetic model (R2: 0.990–0.997 in Figure 1b). The rate constants, ks1, were obtained from the slopes of the straight lines by linear regression of ln(C0/Ct)-time plots. The change of ks1 with fluorination is similar to that of the DR (Table 1). A 2.5-fold increment in ks1 was achieved from Fe3O4-blank to F-Fe3O4-1 (0.0011 vs. 0.0284 min−1) at 40 °C.
Shown in Figure 2a,b is the effect of the initial H2O2 concentration, [H2O2]0, on the decolorization of OG at 40 °C. The decolorization efficiency is lower at lower [H2O2]0—the 2 h DR and ks1 value were only 65.9% and 0.0082 min−1, respectively, at 20 mM of [H2O2]0 in Table S1. On the other hand, higher [H2O2]0 was not beneficial either—the 2 h DE and ks1 value were 90.3% and 0.0182 min−1, respectively, at 80 mM of [H2O2]0. The optimal [H2O2]0 was about 40–60 mM for efficient decolorization of OG (0.1 mM, pH 6.5) on F-Fe3O4-1 at 40 °C (the 2 h DE and ks1 value were ~97% and 0.030 min−1, respectively). The existence of the optimal H2O2 concentration is due to the dual roles of H2O2. Firstly, it is the source of ⋅OH radicals for the OG degradation. Secondly, excess H2O2 scavenges ⋅OH radicals (Equation (2)) [26,27], thus lowering the concentration of ⋅OH radicals in the solution.
H2O2 + ⋅OH → HOO⋅ + H2O
The OG decolorization at different pH values is presented in Figure 2c,d. Both the 2 h DR and ks1 value increased when the initial pH value reduced from 9.0 to 3.0. It is noteworthy that the high decolorization efficiencies of the OG (0.1 mM) solution could be achieved (2 h DR: 94.8–96.8% and ks1: 0.0252–0.0284 min−1 in Table S2) at the natural pH value of the OG (0.1 mM) solution (6.4–6.8). This has technological and operational advantages since it is not required to pre-adjust its pH value.
The temperature effect on the OG decolorization is displayed in Figure 2e,f. It is obvious that the 2 h DE and k1 value increased as the reaction temperature increased from 25 °C to 50 °C (e.g., they were 32.6% and 0.0028 min−1 at 25 °C, and increased to 99.5% and 0.0445 min−1 at 55 °C in Table S3). The kinetic parameters, including the activation energy (Ea) and collision frequency (A, also called pre-exponential factor), of the OG decolorization, were estimated from the linear Arrhenius equation of lnks1 = lnA − Ea/RT, where R is the universal gas constant (8.314 J/(mol⋅K)). The Ea value was estimated to be 75.7 kJ/mol from the slope of the fitted linear equation of the plot of lnks1 − 1000/T (Figure S1), which is significantly lower than the dissociation energy of H2O2 to ⋅OH (213.8 kJ/mol) due to the catalytic effect of F-Fe3O4-1 [28] but higher than those obtained for the RhB degradation on Fe3O4 MNPs (45.6 kJ/mol [26], 46.6 kJ/mol [29], 47.6 kJ/mol [10] or 51 kJ/mol [30]), which may be attributed to the smaller surface area of the F-Fe3O4-1 microspheres. However, it is interesting to note that the pre-exponential factor A in the reaction system of OG/F-Fe3O4-1 microspheres is five orders of magnitude larger than that obtained in RhB/Fe3O4 MNPs (6.88 × 1010 min−1 vs. 8.26 × 105 min−1 [10]), which indicates that collision frequency between reacting species are greatly enhanced on F-Fe3O4-1 microspheres in agreement with the proposed adsorption-enhanced catalytic degradation of anionic molecules (vide infra).
To check the scope of the applicability of F-Fe3O4-1 in organic wastewater treatment, the CWPO degradations of other organic dyes (CR, MB, and RhB) were investigated and displayed in Figure 3 and Table 2. As shown in Figure 3a,b, the degradation of another anion dye CR was also significantly promoted on F-Fe3O4-1 than on Fe3O4-blank (the 2 h DR and ks1 increased from 19.2% and 0.0012 min−1 for Fe3O4-blank to 91.7% and 0.024 min−1 for F-Fe3O4-1). However, the promotional effect of fluorination was not observed for cationic dye MB (2 h DR and ks1: 44.2% and 0.0036 min−1 on Fe3O4-blank versus 42.4% and 0.0040 min−1 on F-Fe3O4-1) and became less obvious for the RhB dye (2 h DR and ks1: 16.4% and 0.0039 min−1 on Fe3O4-blank versus 67.5% and 0.0080 min−1 on F-Fe3O4-1). The pKa value of RhB was reported to be 3.5 [31], 4.1 ± 0.1 [32,33], or 6.41 [27]. Therefore, the carboxylic acid group of RhB is deprotonated and becomes negatively charged at the tested pH (~6.6), which may partially contribute to the enhanced adsorption and degradation of RhB on F-Fe3O4-1.
F-Fe3O4-1 maintained its high catalytic activity after continuously running for 16 h in recyclability tests (Figure 4 and Table 3). The 2 h DR, ks1, and chemical oxygen demand (COD) reduction rate were 84.9%, 0.0151 min−1, and 65.2%, respectively, in the eighth run of F-Fe3O4-1. The low-to-moderate COD reduction rate (65.2–83.2%) indicated the formation of some organic intermediates during the degradation of OG with H2O2. These intermediates may cover some active sites on F-Fe3O4-1, resulting in a gradual reduction in the activity of reused F-Fe3O4-1.

2.2. Characterization of F-Fe3O4-r Microspheres

The F-Fe3O4-r microspheres used in this work have been characterized in our previous work [25]. The XRD patterns shown in Figure 1 of Reference 25 indicate that F-Fe3O4-r (r ≤ 1) consist of the single magnetite (PDF #19-0629) phase while F-Fe3O4-r (r = 2, 3) contain the Na3FeF6 (PDF #22-1381) impurity. Unless otherwise specified, the fluorinated Fe3O4 in this work referred to F-Fe3O4-1 microspheres, which comprise pure magnetite with the highest fluoride content (7.1 ± 1.4 wt% determined by EDX in Table S4) and the best catalytic performance (Figure 1). In comparison with Fe3O4-blank, F-Fe3O4-1 microspheres have a reduced average particle size (360 ± 80 nm vs. 220 ± 34 nm) with slightly more aggregation (Figure 2 in Reference [25]), which resulted in their smaller SBET value (28.0 and 17.6 m2/g) (Figure 3 in Reference [25]). However, there is no obvious change in the magnetic properties of Fe3O4 upon fluorination (Figure 4 in Reference [25]), which is beneficial for the magnetic separation and recycling of the F-Fe3O4-1 microspheres.

2.3. Mechanistic Investigation

The concentrations of dissolved F and iron ions after OG decolorization were, respectively, 0.65 mg/L and 0.15 mg/L, which meet the criteria set in the fourth edition of the World Health Organization’s Guidelines for drinking-water quality (0.5–1 mg/L fluoride and <0.3 mg/L iron). In addition, the decolorization of OG (0.1 mM) was negligible (~5%) after replacing F-Fe3O4-1 with 0.65 mg/L of F or 0.15 mg/L of Fe2+ (or Fe3+) under otherwise identical conditions (Figure 5), which supported the notion that the decolorization of OG with H2O2 was catalyzed by heterogeneous F-Fe3O4-1 solid catalyst, not by soluble species.
The addition of tert-butanol (10 mM) as a scavenger of hydroxyl radicals significantly inhibited the degradation of OG with H2O2 at pH 6.5 over F-Fe3O4-1. As shown in Figure 6 and Entry 6 in Table 1, the 2 h DR and ks1, respectively, decreased from 96.8% and 0.0284 min−1 to 35.2% and 0.0026 min−1, implying that hydroxyl radicals are the main reactive oxygen species contributing to the OG degradation. Since the decolorization of OG was not completely suppressed by tert-butanol (10 mM), other reactive oxygen species such as superoxide radical anions may also make a minor contribution to the decolorization of OG.
The EPR spectra with the DMPO spin-trapping technique were used to identify the generation of ·OH radicals. As shown in Figure 7, a typical 1:2:2:1 quadruplet pattern corresponding to the DMPO-·OH adduct was observed for both F-Fe3O4-1 and Fe3O4-blank, supporting the formation of ·OH radicals in the H2O2/F-Fe3O4-1 (or Fe3O4-blank) reaction system. However, the signal intensities for F-Fe3O4-1 were obviously stronger than those for Fe3O4-blank, indicating the higher activity of the former to activate H2O2 due to the larger number of Fe2+ ions generated after fluorination of F-Fe3O4-1 [25].
The improved catalytic activity of F-Fe3O4-1 for the degradation of anionic dyes (e.g., OG and CR) at pH ~7 was related to their enhanced adsorption. According to the measurement of zeta potentials at different pH values (Figure 8), the IEPs of F-Fe3O4-1 and Fe3O4-blank were determined to be 8.7 and 4.9, respectively. The higher IEP (8.7) of F-Fe3O4-1 was attributed to the partial substitution of F for lattice O2− in Fe3O4. As a result, the surface of F-Fe3O4-1 was positively charged, while Fe3O4-blank was negatively charged at pH ~7. Therefore, the adsorption of anionic dye molecules on F-Fe3O4-1 was facilitated via electrostatic attraction but unfavorable on Fe3O4-blank at pH ~7 due to electrostatic repulsion (e.g., the 15-min adsorption of OG was 5.9% on Fe3O4-blank vs. 12.3% on F-Fe3O4-1). It is generally accepted that surface-generated hydroxyl radicals are highly reactive and short-lived [34] and are mostly consumed on the surface before diffusing to the solution [35]. Therefore, the adsorbed anionic molecules (OG or CR), not those in the solution, were more likely to be oxidized by surface-generated hydroxyl radicals. This strengthens the critical role of the adsorption of the solid surface in determining the oxidation rate of pollutants in iron oxide/H2O2 systems [36].

3. Materials and Methods

3.1. Chemicals and Regents

All chemicals, including FeCl3·6H2O, NaOAc, NaF, ethylene glycol, orange G (OG), congo Red (CR), methylene blue (MB), rhodamine B (RhB), and 5,5-dimethyl-1-pyrroline N-oxide (DMPO), were of analytic grade or above and used as received. All aqueous solutions were prepared with deionized (DI) water (<20 μS/cm) produced from Heal Force NW 15VF.

3.2. Synthesis of Fluoride-Modified Fe3O4 Microspheres

The fluoride-modified Fe3O4 microspheres were made via glycothermal synthesis according to the reported method [25] and named as F-Fe3O4-r, where r indicates the nominal F/Fe molar ratio (r: 0–3). The product made at r = 0 was unmodified (denoted Fe3O4-blank). In brief, NaOAc (22 mmol) and NaF (0–21 mmol) were sequentially added at an interval of 10 min to 4.5 mL of FeCl3·6H2O (1.5 M) solution in ethylene glycol and then heated in a 60-mL Teflon-lined stainless steel autoclave at 198 °C for 14 h to obtain F-Fe3O4-r black particles.

3.3. CWPO of Organic Dyes

A general procedure of CWPO was as follows: A mixture of 0.025 g of F-Fe3O4-r microspheres and 50 mL of OG (Cs = 0.1 mM) stock solution was first stirred at 750 rpm and 40 ± 2 °C for 15 min to reach the adsorption equilibrium. The concentration of OG was then measured and taken as the concentration at time zero (C0). The degradation of OG was initiated by rapidly adding 0.230 g of H2O2 (30%) into the mixture. At selected intervals, about 3 mL of the samples was taken by a syringe and passed through a filter membrane, and the residual concentration of OG in solution, Ct, was analyzed. Multiple measurements of the catalytic activity were performed with the relative standard deviation usually below 5%. In recyclability tests, the used F-Fe3O4-1 was magnetically recovered and directly reused for the next run according to the same procedure. This procedure was also applied to degrade CR, MB, and RhB. The concentrations, Ct, of OG, CR, MB, and RhB in aqueous solutions were analyzed by measuring the absorbance at 478, 495, 664, and 554 nm, respectively, using a UV-vis spectrophotometer (UV-1800PC, Shanghai Mapada Instrument, Shanghai, China). Before measurement, the pH value of the CR solution in the quartz cuvette was adjusted by adding one drop of 2 wt% KOH.

3.4. Physicochemical Characterization

Powder X-ray diffraction (XRD) patterns were collected on Bruker D8 Advance Diffractometer with Cu Kα radiation at 30 kV and 20 mA. Scanning electron microscope (SEM) images were taken on JEOL JSM-5510LV SEM (Tokyo, Japan) at 20 kV. N2 adsorption–desorption isotherms were measured on Micromeritics ASAP 2020 (Georgia, United States) at 77 K. The samples were degassed under vacuum at 100 °C for 2 h before measurements. The BET specific surface area SBET was calculated from the adsorbed quantities in the relative pressure (P/P0) range of 0.05–0.30. M-H curves were measured on a JDAW-2000D (Jilin, China) vibrating sample magnetometer (VSM) at room temperature. Zeta potentials were determined at different pH values on Malvern Zetasizer Zen3690 (Malvern, United Kingdom). Electron-spin resonance (EPR) spectra were collected on Bruker EMXplus (Berlin, Germany) at room temperature. The concentrations of fluoride and iron ions in the solution were analyzed using ionic chromatography (ICS-1500) and atomic absorption spectroscopy (Agilent 240 AA, California, United States), respectively.

4. Conclusions

Fluorination increased the isoelectric point from 4.9 for undoped Fe3O4-blank microspheres to 8.7 for F-Fe3O4-1 with 7.1 ± 1.4 wt% of fluoride. In comparison with Fe3O4-blank, F-Fe3O4-1 could efficiently catalyze the decolorization of anionic dye solutions at pH ~7. This was mainly attributed to the enhanced adsorption of anionic dyes on the surface of F-Fe3O4-1 via electrostatic attraction, followed by oxidative degradation via a free-radical reaction pathway. This work strengthens the critical role of the adsorption of the solid surface in determining the oxidation rate of pollutants in iron oxide/H2O2 systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12121564/s1, Figure S1: The Arrhenius plot of the degradation of OG (0.1 mM, pH 6.5) with H2O2 on F-Fe3O4-1 microspheres; Table S1: Effect of initial H2O2 concentration on OG decolorization with H2O2 over F-Fe3O4-1 microspheres; Table S2: Effect of initial pH on OG decolorization with H2O2 over F-Fe3O4-1 microspheres; Table S3: Effect of reaction temperature on OG decolorization with H2O2 over F-Fe3O4-1 microspheres; and Table S4: EDX-determined elemental composition of F-Fe3O4-1 microspheres.

Author Contributions

Conceptualization, F.C.; methodology, investigation, data curation, and formal analysis, H.L., W.C. and F.C.; resources, F.C. and R.C.; and writing—original draft preparation, review, and editing, F.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Natural Science Foundation of China (No. 21571146).

Data Availability Statement

Data are available within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ribeiro, R.S.; Silva, A.M.T.; Figueiredo, J.L.; Faria, J.L.; Gomes, H.T. Catalytic wet peroxide oxidation: A route towards the application of hybrid magnetic carbon nanocomposites for the degradation of organic pollutants. A review. Appl. Catal. B Environ. 2016, 187, 428–460. [Google Scholar] [CrossRef]
  2. Munoz, M.; de Pedro, Z.M.; Casas, J.A.; Rodriguez, J.J. Preparation of magnetite-based catalysts and their application in heterogeneous Fenton oxidation—A review. Appl. Catal. B Environ. 2015, 176–177, 249–265. [Google Scholar] [CrossRef] [Green Version]
  3. Huang, R.; Fang, Z.; Yan, X.; Cheng, W. Heterogeneous sono-Fenton catalytic degradation of bisphenol A by Fe3O4 magnetic nanoparticles under neutral condition. Chem. Eng. J. 2012, 197, 242–249. [Google Scholar] [CrossRef]
  4. Pham, A.L.-T.; Lee, C.; Doyle, F.M.; Sedlak, D.L. A silica-supported iron oxide catalyst capable of activating hydrogen peroxide at neutral pH values. Environ. Sci. Technol. 2009, 43, 8930–8935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Zhu, Y.; Zhu, R.; Xi, Y.; Zhu, J.; Zhu, G.; He, H. Strategies for enhancing the heterogeneous Fenton catalytic reactivity: A review. Appl. Catal. B Environ. 2019, 255, 117739. [Google Scholar] [CrossRef]
  6. Pouran, S.R.; Raman, A.A.A.; Daud, W.M.A.W. Review on the application of modified iron oxides as heterogeneous catalysts in Fenton reactions. J. Clean. Prod. 2014, 64, 24–35. [Google Scholar] [CrossRef] [Green Version]
  7. Bokare, A.D.; Choi, W. Review of iron-free Fenton-like systems for activating H2O2 in advanced oxidation processes. J. Hazard. Mater. 2014, 275, 121–135. [Google Scholar] [CrossRef]
  8. Yin, Y.; Shi, L.; Li, W.; Li, X.; Wu, H.; Ao, Z.; Tian, W.; Liu, S.; Wang, S.; Sun, H. Boosting Fenton-like reactions via single atom Fe catalysis. Environ. Sci. Technol. 2019, 53, 11391–11400. [Google Scholar] [CrossRef]
  9. Gao, L.; Zhuang, J.; Nie, L.; Zhang, J.; Zhang, Y.; Gu, N.; Wang, T.; Feng, J.; Yang, D.; Perrett, S.; et al. Intrinsic peroxidase-like activity of ferromagnetic nanoparticles. Nat. Nanotechnol. 2007, 2, 577–583. [Google Scholar] [CrossRef]
  10. Chen, F.; Xie, S.; Huang, X.; Qiu, X. Ionothermal synthesis of Fe3O4 magnetic nanoparticles as efficient heterogeneous Fenton-like catalysts for degradation of organic pollutants with H2O2. J. Hazard. Mater. 2017, 322, 152–162. [Google Scholar] [CrossRef]
  11. Huang, X.; Xu, C.; Ma, J.; Chen, F. Ionothermal synthesis of Cu-doped Fe3O4 magnetic nanoparticles with enhanced peroxidase-like activity for organic wastewater treatment. Adv. Powder Technol. 2018, 29, 796–803. [Google Scholar] [CrossRef]
  12. Xu, C.; Zhang, J.; Huang, X.; Chen, F. One-step nonaqueous synthesis of modified magnetic Fe3O4 microspheres by using epichlorohydrin as functional solvent. Chem. Lett. 2019, 48, 22–25. [Google Scholar] [CrossRef]
  13. Huang, X.; Du, W.; Chen, R.; Chen, F. Adsorption-enhanced catalytic wet peroxide oxidation of aromatic compounds on ionothermally synthesised copper-doped magnetite magnetic nanoparticles. Environ. Chem. 2020, 17, 426–435. [Google Scholar] [CrossRef]
  14. Liang, X.; Zhong, Y.; He, H.; Yuan, P.; Zhu, J.; Zhu, S.; Jiang, Z. The application of chromium substituted magnetite as heterogeneous Fenton catalyst for the degradation of aqueous cationic and anionic dyes. Chem. Eng. J. 2012, 191, 177–184. [Google Scholar] [CrossRef]
  15. Wang, J.; Zheng, S.; Shao, Y.; Liu, J.; Xu, Z.; Zhu, D. Amino-functionalized Fe3O4@SiO2 core–shell magnetic nanomaterial as a novel adsorbent for aqueous heavy metals removal. J. Colloid Interf. Sci. 2010, 349, 293–299. [Google Scholar] [CrossRef]
  16. Cai, H.; Li, X.; Ma, D.; Feng, Q.; Wang, D.; Liu, Z.; Wei, X.; Chen, K.; Lin, H.; Qin, S.; et al. Stable Fe3O4 submicrospheres with SiO2 coating for heterogeneous Fenton-like reaction at alkaline condition. Sci. Total Environ. 2021, 764, 14420. [Google Scholar] [CrossRef]
  17. Zhang, J.; Wang, Z.; Chen, R.; Chen, F. New soft chemistry route to titanomagnetite magnetic nanoparticles with enhanced peroxidase-like activity. Powder Technol. 2020, 373, 39–45. [Google Scholar] [CrossRef]
  18. Du, W.; Huang, R.; Huang, X.; Chen, R.; Chen, F. Copper-promoted heterogeneous Fenton-like oxidation of Rhodamine B over Fe3O4 magnetic nanocatalysts at mild conditions. Environ. Sci. Pollut. Res. 2021, 28, 19959–19968. [Google Scholar] [CrossRef]
  19. Yu, J.C.; Yu, J.; Ho, W.K.; Jiang, Z.; Zhang, L. Effects of F-doping on the photocatalytic activity and microstructures of nanocrystalline TiO2 powders. Chem. Mater. 2002, 14, 3808–3816. [Google Scholar] [CrossRef]
  20. Kumar, V.; Govind, A.; Nagarajan, R. Optical and photocatalytic properties of heavily F-doped SnO2 nanocrystals by a novel single-source precursor approach. Inorg. Chem. 2011, 50, 5637–5645. [Google Scholar] [CrossRef]
  21. Gasparotto, A.; Barreca, D.; Bekermann, D.; Devi, A.; Fischer, R.A.; Fornasiero, P.; Gombac, V.; Lebedev, O.I.; Maccato, C.; Montini, T.; et al. F-doped Co3O4 photocatalysts for sustainable H2 generation from water/ethanol. J. Am. Chem. Soc. 2011, 133, 19362–19365. [Google Scholar] [CrossRef] [PubMed]
  22. Zhu, L.-P.; Wang, L.-L.; Bing, N.-C.; Huang, C.; Wang, L.-J.; Liao, G.-H. Porous fluorine-doped γ-Fe2O3 hollow spheres: Synthesis, growth mechanism, and their application in photocatalysis. ACS Appl. Mater. Interfaces 2013, 5, 12478–12487. [Google Scholar] [CrossRef] [PubMed]
  23. Yu, L.; Xiong, L.; Yu, Y. Cu2O homojunction solar cells: F-doped n-type thin film and highly improved efficiency. J. Phys. Chem. C 2015, 119, 22803–22811. [Google Scholar] [CrossRef]
  24. Fu, H.; Zhang, S.; Xu, T.; Zhu, Y.; Chen, J. Photocatalytic degradation of RhB by fluorinated Bi2WO6 and distributions of the intermediate products. Environ. Sci. Technol. 2008, 42, 2085–2091. [Google Scholar] [CrossRef]
  25. Chen, W.; Zhang, J.; Chen, F. Glycothermal synthesis of fluorinated Fe3O4 microspheres with distinct peroxidase-like activity. Adv. Powder Technol. 2019, 30, 999–1005. [Google Scholar] [CrossRef]
  26. Luo, W.; Zhu, L.; Wang, N.; Tang, H.; Cao, M.; She, Y. Efficient removal of organic pollutants with magnetic nanoscaled BiFeO3 as a reusable heterogeneous Fenton-like catalyst. Environ. Sci. Technol. 2010, 44, 1786–1791. [Google Scholar] [CrossRef]
  27. Xue, X.; Hanna, K.; Deng, N. Fenton-like oxidation of Rhodamine B in the presence of two types of iron (II, III) oxide. J. Hazard. Mater. 2009, 166, 407–414. [Google Scholar] [CrossRef]
  28. Dean, J.A. Lange’s Handbook of Chemistry, 15th ed.; McGraw-Hill: New York, NY, USA, 1998. [Google Scholar]
  29. Wang, N.; Zhu, L.; Wang, M.; Wang, D.; Tang, H. Sono-enhanced degradation of dye pollutants with the use of H2O2 activated by Fe3O4 magnetic nanoparticles as peroxidase mimetic. Ultrason. Sonochem. 2010, 17, 78–83. [Google Scholar] [CrossRef]
  30. Lin, C.-C.; Ho, J.-M.; Hsieh, H.-L. Feasibility of using a rotating packed bed in preparing Fe3O4 nanoparticles. Chem. Eng. J. 2012, 203, 88–94. [Google Scholar] [CrossRef]
  31. Anandkumar, J.; Mandal, B. Adsorption of chromium(VI) and Rhodamine B by surface modified tannery waste: Kinetic, mechanistic and thermodynamic studies. J. Hazard. Mater. 2011, 186, 1088–1096. [Google Scholar] [CrossRef]
  32. Chen, F.; Zhao, J.; Hidaka, H. Highly selective deethylation of rhodamine B: Adsorption and photooxidation pathways of the dye on the TiO2/SiO2 composite photocatalyst. Int. J. Photoenergy 2003, 5, 209–217. [Google Scholar] [CrossRef] [Green Version]
  33. Guo, Y.; Zhao, J.; Zhang, H.; Yang, S.; Qi, J.; Wang, Z.; Xu, H. Use of rice husk-based porous carbon for adsorption of Rhodamine B from aqueous solutions. Dye. Pigment. 2005, 66, 123–128. [Google Scholar] [CrossRef]
  34. Guo, S.; Wang, H.; Yang, W.; Fida, H.; You, L.; Zhou, K. Scalable synthesis of Ca-doped α-Fe2O3 with abundant oxygen vacancies for enhanced degradation of organic pollutants through peroxymonosulfate activation. Appl. Catal. B Environ. 2020, 262, 118250. [Google Scholar] [CrossRef]
  35. Lin, S.-S.; Gurol, M.D. Catalytic decomposition of hydrogen peroxide on iron oxide: Kinetics, mechanism, and implications. Environ. Sci. Technol. 1998, 32, 1417–1423. [Google Scholar] [CrossRef]
  36. Hanna, K.; Kone, T.; Medjahdi, G. Synthesis of the mixed oxides of iron and quartz and their catalytic activities for the Fenton-like oxidation. Catal. Commun. 2008, 9, 955–959. [Google Scholar] [CrossRef]
Scheme 1. Molecular structures of representative organic dyes used in this work.
Scheme 1. Molecular structures of representative organic dyes used in this work.
Catalysts 12 01564 sch001
Figure 1. Catalytic degradation of OG with H2O2 over F-Fe3O4-r (r = 0–3). (a) Ct/Cs-time plots and (b) ln(C0/Ct)-time plots. Reaction conditions: OG (0.1 mM, pH 6.8), H2O2: 40 mM, and catalyst loading: 0.5 g/L at 40 °C.
Figure 1. Catalytic degradation of OG with H2O2 over F-Fe3O4-r (r = 0–3). (a) Ct/Cs-time plots and (b) ln(C0/Ct)-time plots. Reaction conditions: OG (0.1 mM, pH 6.8), H2O2: 40 mM, and catalyst loading: 0.5 g/L at 40 °C.
Catalysts 12 01564 g001
Figure 2. Effects of reaction parameters on OG decolorization with H2O2 over F-Fe3O4-1. (a,c,e) Ct/Cs-time plots and (b,d,f) ln(C0/Ct)-time plots. General reaction conditions: OG (0.10 mM, pH 6.5), H2O2: 40 mM, and F-Fe3O4-1: 0.5 g/L at 40 °C for 2 h unless otherwise specified in the legends.
Figure 2. Effects of reaction parameters on OG decolorization with H2O2 over F-Fe3O4-1. (a,c,e) Ct/Cs-time plots and (b,d,f) ln(C0/Ct)-time plots. General reaction conditions: OG (0.10 mM, pH 6.5), H2O2: 40 mM, and F-Fe3O4-1: 0.5 g/L at 40 °C for 2 h unless otherwise specified in the legends.
Catalysts 12 01564 g002
Figure 3. CWPO of organic dyes over F-Fe3O4-r (r = 0 and 1) microspheres. (a,c,e) Ct/Cs-time plots and (b,d,f) ln(C0/Ct)-time plots. Reaction conditions: CR (100 mg/L, pH 6.8) (MB (0.01 mM, 6.5) or RhB (0.01 mM, 6.6)), H2O2: 40 mM, catalyst loading: 0.5 g/L at 25 °C except 40 °C for CR.
Figure 3. CWPO of organic dyes over F-Fe3O4-r (r = 0 and 1) microspheres. (a,c,e) Ct/Cs-time plots and (b,d,f) ln(C0/Ct)-time plots. Reaction conditions: CR (100 mg/L, pH 6.8) (MB (0.01 mM, 6.5) or RhB (0.01 mM, 6.6)), H2O2: 40 mM, catalyst loading: 0.5 g/L at 25 °C except 40 °C for CR.
Catalysts 12 01564 g003aCatalysts 12 01564 g003b
Figure 4. Recyclability tests of F-Fe3O4-1 for catalytic degradation of OG with H2O2. (a) Ct/Cs-time plots and (b) ln(C0/Ct)-time plots. Reaction conditions: fresh F-Fe3O4-1: 0.5 g/L, OG (0.1 mM, pH 6.5), and H2O2: 40 mM at 40 °C for 2 h. The used F-Fe3O4-1 was magnetically recovered and directly re-used as the fresh.
Figure 4. Recyclability tests of F-Fe3O4-1 for catalytic degradation of OG with H2O2. (a) Ct/Cs-time plots and (b) ln(C0/Ct)-time plots. Reaction conditions: fresh F-Fe3O4-1: 0.5 g/L, OG (0.1 mM, pH 6.5), and H2O2: 40 mM at 40 °C for 2 h. The used F-Fe3O4-1 was magnetically recovered and directly re-used as the fresh.
Catalysts 12 01564 g004
Figure 5. Catalytic degradation of OG (0.1 mM, pH 6.5) with H2O2 (40 mM) after replacing 0.5 g/L of F-Fe3O4-1 with 0.65 mg/L of F or 0.15 mg/L of Fe2+ (or Fe3+).
Figure 5. Catalytic degradation of OG (0.1 mM, pH 6.5) with H2O2 (40 mM) after replacing 0.5 g/L of F-Fe3O4-1 with 0.65 mg/L of F or 0.15 mg/L of Fe2+ (or Fe3+).
Catalysts 12 01564 g005
Figure 6. Effect of tert-butanol on catalytic degradation of OG with H2O2 over F-Fe3O4-1. (a) C/Cs-time plots and (b) ln(C0/C)-time plots. Reaction conditions: OG (0.1 mM, pH 6.5), H2O2 (40 mM), and F-Fe3O4-1 (0.5 g/L) at 40 °C.
Figure 6. Effect of tert-butanol on catalytic degradation of OG with H2O2 over F-Fe3O4-1. (a) C/Cs-time plots and (b) ln(C0/C)-time plots. Reaction conditions: OG (0.1 mM, pH 6.5), H2O2 (40 mM), and F-Fe3O4-1 (0.5 g/L) at 40 °C.
Catalysts 12 01564 g006
Figure 7. DMPO-trapped EPR spectra. Conditions: F-Fe3O4-1 or Fe3O4-blank (0.5 g/L), H2O2 (40 mM), and DMPO (ca. 20 mM).
Figure 7. DMPO-trapped EPR spectra. Conditions: F-Fe3O4-1 or Fe3O4-blank (0.5 g/L), H2O2 (40 mM), and DMPO (ca. 20 mM).
Catalysts 12 01564 g007
Figure 8. Zeta potentials at different pH values.
Figure 8. Zeta potentials at different pH values.
Catalysts 12 01564 g008
Table 1. Effect of fluorination on catalytic performances of F-Fe3O4-r microspheres for the OG decolorization 1.
Table 1. Effect of fluorination on catalytic performances of F-Fe3O4-r microspheres for the OG decolorization 1.
Entryr15-min Adsorption (%)2-h
DR (%)
ks1/min−1
(R2)
105.917.60.0011 (0.997)
20.110.862.20.0070 (0.990)
3112.396.80.0284 (0.991)
429.868.40.0092 (0.994)
537.847.50.0045 (0.991)
6 218.335.20.0026 (0.996)
1 Reaction reactions were the same as Figure 1; 2 Tert-butanol (10 mM) added.
Table 2. CWPO of organic dyes over F-Fe3O4-r microspheres.
Table 2. CWPO of organic dyes over F-Fe3O4-r microspheres.
Organic Dyesr15-min
Adsorption (%)
Degradation Rate (%)ks1/min−1 (R2)
CR (100 mg/L, pH 6.8)05.219.20.0012 (0.975)
134.391.70.024 (0.976)
MB (0.01 mM, pH 6.5)011.644.20.0036 (0.989)
17.342.40.0040 (0.990)
RhB (0.01 mM, pH 6.6)017.449.30.0039 (0.984)
116.667.10.0080 (0.993)
Table 3. Recyclability of F-Fe3O4-1 for catalytic degradation of OG with H2O2 1.
Table 3. Recyclability of F-Fe3O4-1 for catalytic degradation of OG with H2O2 1.
Run2-h DR (%)ks1/min−1 (R2)COD Reduction Rate (%)
196.80.0284 (0.988)83.2
292.70.0191 (0.978)80.1
389.60.0178 (0.990)75.5
489.10.0170 (0.988)70.4
587.10.0163 (0.985)68.6
686.40.0160 (0.990)66.4
785.90.0157 (0.987)66.1
884.90.0151 (0.987)65.2
1 Reaction conditions were the same as Figure 4.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, F.; Lv, H.; Chen, W.; Chen, R. Catalytic Wet Peroxide Oxidation of Anionic Pollutants over Fluorinated Fe3O4 Microspheres at Circumneutral pH Values. Catalysts 2022, 12, 1564. https://doi.org/10.3390/catal12121564

AMA Style

Chen F, Lv H, Chen W, Chen R. Catalytic Wet Peroxide Oxidation of Anionic Pollutants over Fluorinated Fe3O4 Microspheres at Circumneutral pH Values. Catalysts. 2022; 12(12):1564. https://doi.org/10.3390/catal12121564

Chicago/Turabian Style

Chen, Fengxi, Huaixiang Lv, Wu Chen, and Rong Chen. 2022. "Catalytic Wet Peroxide Oxidation of Anionic Pollutants over Fluorinated Fe3O4 Microspheres at Circumneutral pH Values" Catalysts 12, no. 12: 1564. https://doi.org/10.3390/catal12121564

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