Next Article in Journal
UX in AR-Supported Industrial Human–Robot Collaborative Tasks: A Systematic Review
Previous Article in Journal
Analysis and Optimization of the Robotic Mobile Fulfillment Systems Considering Congestion
Previous Article in Special Issue
Potassium Recovery from Potassium Solution and Seawater Using Different Adsorbents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Degradation of Oxytetracycline by Persulfate Activation Using a Magnetic Separable Iron Oxide Catalyst Derived from Hand-Warmer Waste

1
Department of Environmental and Safety Engineering, Ajou University, Suwon 16499, Korea
2
Department of Bioresources and Rural Systems Engineering, Hankyong National University, Anseong 17579, Korea
3
Department of Agricultural and Biological Chemistry, Chonnam National University, Gwangju 61186, Korea
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2021, 11(21), 10447; https://doi.org/10.3390/app112110447
Submission received: 4 October 2021 / Revised: 1 November 2021 / Accepted: 5 November 2021 / Published: 7 November 2021

Abstract

:
Oxytetracycline (OTC) is a tetracycline antibiotic that is widely used in the drug therapy and livestock industry and may threaten human health and ecosystems when released into the environment. In this study, a catalyst was prepared from hand-warmer waste using a simple magnetic separation method. The prepared hand-warmer waste catalyst (HWWC) was used as a persulfate (PS) activator for OTC removal. Characterization methods, such as X-ray diffraction and scanning electron microscopy–energy dispersive X-ray spectrometry, were used to investigate the crystal structure, surface morphology, and weight ratios of the elements in the HWWC. The degradation efficiency of OTC in the presence of the catalyst and PS was studied, and the radical generation mechanism of the catalyst was investigated. The removal ratio of OTC by PS activation was greater than 99% for a reaction time of 24 min at a pH of 6. The effects of the HWWC dosage, PS concentration, and solution pH on OTC degradation were also investigated. The reuse test revealed that HWWC can be reused for eight cycles with great stability. These results suggest that PS activation using hand-warmer waste can be an efficient strategy for the degradation of antibiotics.

1. Introduction

The amount of antibiotics released into the environment is increasing owing to the lack of appropriate disposal methods and strict control measures, which threaten human health and the ecosystem [1,2]. Oxytetracycline (OTC), an antibiotic, is widely used as an antimicrobial agent and growth factor in drug therapy and the livestock industry [3,4]. Approximately 70% of OTC leaves organisms via urine and feces without undergoing metabolism because of its poor absorption [5]. Therefore, OTC has been detected in various environments, such as aquatic systems, soil, and sediments [6,7,8]. In aquatic environments, several studies have reported that OTC has been detected in river water [9] and in the influent and effluent of a wastewater treatment plant [10,11,12,13]. However, it is difficult to degrade the released OTC in water using conventional wastewater treatment processes [14]. Therefore, effective treatment of OTC in water is a problem that needs to be solved urgently.
Sulfate radicals (SO4) and hydroxyl radicals (HO•) are widely used reactive radical species in wastewater treatment because of their high oxidizing capabilities [15]. Generally, activating peroxides such as persulfate (PS), peroxymonosulfate, and hydrogen peroxide or photocatalytic processes can generate these radical species [16]. Among the abovementioned peroxides, PS is much cheaper and easier to activate owing to its low band energy (140 kJ/mol) [17,18], PS has attracted attention as an oxidant for degrading various pollutants [19,20]. Catalysts such as metal-containing oxides and transition metals have been used to activate PS because they are energy-free and economic [21,22,23]. Moreover, research on the reuse of waste containing metal elements such as Fe as a PS activator has been conducted [16,24,25].
Disposable hand warmers are widely used to keep oneself warm; thus, the demand for hand warmers greatly increases in winter. After exposure to air, the materials in the hand warmer pocket react and release heat for a period of time. The spent hand warmer is then discarded, which can adversely affect the environment and lead to wastage of resources [26], recycling or reusing the spent hand warmer is needed to reduce environmental pollution. Hand-warmer waste generally contains iron oxide (Fe2O3) particles. Therefore, reusing hand-warmer waste for the activation of PS can be an environmentally friendly and cost-saving technique. To the best of our knowledge, this study is the first to recycle hand-warmer waste as a catalyst for PS activation.
In this study, a hand-warmer waste catalyst (HWWC) was prepared by a simple magnetic separation method and used as a PS activator for OTC degradation. The surface morphology and crystal structure of the prepared HWWC were investigated. The effects of the catalyst dosage, PS concentration, and pH on the degradation of OTC were studied. In addition, the stability of the catalyst was evaluated by conducting a reuse test.

2. Materials and Methods

2.1. Chemical and Materials

A hand warmer was obtained from DABONG Industrial Co., Ltd. (Seoul, Korea). Oxytetracycline hydrochloride (C22H24N2O9∙HCl ≥ 97.5%) was purchased from Sigma-Aldrich Co., Ltd. (Burlington, MA, USA). Sodium phosphate monobasic anhydrous (NaH2PO4 ≥ 98%), sodium phosphate dibasic anhydrous (Na2HPO4 ≥ 99.0%), sodium hydroxide (NaOH ≥ 98.0%), hydrogen chloride (HCl ≥ 35.0–37.0%), and acetonitrile (ACN) (CH3CN ≥ 99.9%) were purchased from Samchun Pure Chemical Co., Ltd. (Pyeongtaek-si, Korea). Sodium persulfate (Na2S2O8 ≥ 98%) was purchased from Junsei Chemical Co., Ltd. (Tokyo, Japan). Deionized (DI) water with a resistivity of 18.2 MΩ/cm (Millipore, Darmstadt, Germany) was used to prepare the solutions.

2.2. Catalyst Preparation

HWWC was prepared using a simple magnetic separation method. After a disposable hand warmer was exposed to air for 36 h, 10 g of the contents inside the hand warmer were placed in 1 L of DI water. The Fe2O3 in the DI water was then magnetically separated. The separation process was repeated three times, and the obtained solid was dried in an oven at 80 °C for 24 h. The dried solid was ground for further experiments.

2.3. Experimental Procedure

The OTC degradation experiments were initiated by adding 1 mM PS to 50 mL of the solution containing 20 μM OTC and 0.2 g/L HWWC. The reaction was conducted in a shaking incubator at 150 rpm and 25 °C. The pH of the solution was adjusted to 3, 4, 6, and 8 using 0.1 M NaOH and 0.1 M HCl and analyzed using a pH meter (Orion Star A211, Thermo, Waltham, MA, USA). To perform the reuse test, the catalyst was magnetically separated after each reaction cycle.

2.4. Analytical Method

The OTC concentration was measured using a YL 9100 HPLC system (Youngin Chromass, Anyang, Korea) with a YL 9120 UV/Vis detector and YL 9150 autosampler. A YL C18-4D column (4.6 mm × 150 mm, 5 μm) was used to separate methanol, ACN, and 10 mM phosphate buffer (pH of 7) (15:15:70). The mobile phase was isocratically eluted at a flow rate of 1.0 mL/min. The column temperature was 35 °C, and OTC was detected at 260 nm.

2.5. Characterization

The surface morphology and elemental contents of the HWWC were observed using a scanning electron microscope-energy dispersive X-ray spectrometer (SEM-EDS) (JSM-7900F, JEOL, Tokyo, Japan). The X-ray diffraction (XRD) pattern of the catalyst was analyzed using an XRD system (D/max-2500V, Rigaku, Tokyo, Japan). The point of zero charge (pHpzc) of HWWC was determined by titration method with slight modification [27]. HWWC (0.04 g) was suspended in 20 mL of 0.01 M NaNO3 for 24 h. Then the pH of solution was adjusted using 0.1 M HNO3 or NaOH solution. To reach the equilibrium, the solution was agitated for 1 h, then the pHinitial was measured. After measuring the pHinitial, 0.6 g of NaNO3 was added to the suspension. After 3 h, the pHfinal of the solution was measured. The pHpzc value was determined as ΔpH (pHfinal–pHinitial) was 0 when plotting ΔpH against pHfinal. As shown in Figure 1a, pHpzc of HWWC was 7.4. The magnetic property of HWWC was measured using vibrating sample magnetometer (VSM) (Model 7404, Lake shore cryotronics, Westerville, OH, USA).

3. Results and Discussion

3.1. Crystal Structure, Surface Morphology, and Magnetic Properties of HWWC

Figure 1b shows the XRD patterns of the prepared HWWC. The strongest peak observed at approximately 2θ = 35.58° indicated a reduction in the (119) diffraction of γ-Fe2O3 [28]. The peaks observed at approximately 30.08°, 43.16°, 57.16°, and 62.80° corresponding to (205), (0012), (1115), and (4012) were also in good agreement with those of γ-Fe2O3 [28,29]. Other impurity peaks at approximately 33.30°, 54.10°, and 63.40° agreed with α-Fe2O3 [30]. These results indicated that the magnetically separated HWWC was a mixture mainly consisting of γ-Fe2O3 and α-Fe2O.
Figure 2 shows the SEM images and EDS spectra of the HWWC particles. The morphology of the particles was approximately spherical (Figure 2a). Based on the elemental analysis (O = 32.23%, Fe = 66.98%) (Figure 2b), it was determined that the sample mainly consisted of Fe and O, therefore confirming that the produced powder was Fe2O3, which agreed with the XRD results.
Figure 3 shows the VSM analysis result of the HWWC particles. The saturation magnetization of the HWWC was determined to be 34.14 emu/g, which was sufficient (>16.3 emu/g) for it to be magnetically recovered from solution using a conventional magnet [31,32]. Thus, HWWC can be easily recovered from water through magnetic separation and reused.

3.2. Control Experiments

Figure 4 shows the removal of OTC under different experimental conditions. OTC was removed when both PS and HWWC were present. The degradation efficiency of OTC by PS activation was greater than 99% in 24 min, and the estimated pseudo first-order rate constant (k) was 0.21 ± 0.03 min−1. This removal rate was comparable to OTC degradation through the Fenton process using H2O2/Fe2+ (kapp = 0.068–0.213 min−1) [33], which indicates that the PS activation process using HWWC can be a promising technique for removing antibiotics from water. The removal rate of OTC in the presence of PS and HWWC in 24 min was low (<6.0%). The degradation mechanism of OTC by PS activation can be expressed by the following equations (Equations (1)–(5)) [34]:
Pollutant + Fe(III) → Pollutant + Fe(II)
Fe(II) + S2O82− → Fe(III) + SO4•− + SO42−
Fe(III) + H2O → Fe(II) + HO + H+
H2O + SO4•− → H+ + HO + SO42−
SO4•− + HO + Pollutant → CO2 + H2O
The electrons could be transferred to Fe(III) when the pollutant was adsorbed onto the Fe2O3 surface (Equation (1)). Therefore, a Fenton-like reaction occurred between S2O82− and Fe(II) at the surface of Fe2O3, therefore generating SO4•− and reforming Fe(III) (Equation (2)). HO also might have been formed by this reaction and contributed to pollutant degradation (Equations (3)–(5)) [35,36,37]. Therefore, the pollutant could be degraded by the generated surface-adsorbed radicals (SO4•− and HO) that might diffuse into the aqueous solution (Equation (5)) [34].
Figure 5 shows the effect of two radical scavengers, ethanol (EtOH) and t-butanol (TBA), on the degradation of OTC. According to equations (Equations (6)–(9)) [38], the reaction rate of EtOH and HO is 50 times faster than that of EtOH and SO4•−, while TBA reacts with HO almost 1000 times faster than that with SO4•−. From the results, the oxidation of OTC was significantly reduced with the addition of TBA and EtOH, suggesting that EtOH can effectively inhibit the oxidation efficiency. In addition, compared with EtOH, the removal rate of OTC was found to be lower in the presence of TBA, which demonstrates that a small amount of SO4•− remaining in the solution can still decompose OTC. Thus, it can be concluded that SO4•− is the dominant oxidizing species in the OTC degradation process by persulfate activation using HWWC catalysts.
EtOH + HO → Intermediates (k = (1.2–2.8) × 109 M−1s−1)
EtOH + SO4•− → Intermediates (k = (1.6–7.7) × 107 M−1s−1)
TBA + HO → Intermediates (k = (3.8–7.6) × 108 M−1s−1)
TBA + SO4•− → Intermediates (k = (4.0–9.1) × 105 M−1s−1)

3.3. Effects of HWWC Dosage and PS Concentration on OTC Degradation

The catalyst dosage and PS concentration are important factors in pollutant degradation. Therefore, degradation experiments with various HWWC dosages (0.05, 0.20, and 0.40 g/L) and PS concentrations (0.5, 1.0, and 2.0 mM) were investigated. As shown in Figure 6, the degradation efficiency of OTC increased in proportion to the HWWC dosage and PS concentration. The k value increased from 0.02 ± 0.00 min−1 to 0.14 ± 0.02 min−1 and increased from 0.09 ± 0.01 min−1 to 0.12 ± 0.01 min−1 when the HWWC dosage and PS concentration increased, respectively. HWWC and PS were the activator and source of reactive radical species, respectively. Thus, their increase could promote OTC degradation [39].

3.4. Effects of pH on OTC Degradation

In heterogeneous PS activation, the initial pH has a significant influence on the degradation of pollutants [40]. Therefore, the degradation efficiency at various solution pH conditions (3, 4, 6, and 8) was investigated (Figure 7). As shown in Figure 5, the removal ratio of OTC at 4 min was 70.8 ± 7.0%, 70.7 ± 0.8%, 31.0 ± 1.9%, and 28.6 ± 0.9% at a pH of 3, 4, 6, and 8, respectively. The degradation performance at a pH of 3 and 4 showed no significant change, but the efficiency significantly decreased when the solution pH increased from 4 to 8, therefore implying that acidic conditions were more favorable for the degradation of OTC by heterogeneous PS activation. This result was observed because the surface properties of the catalyst and the lifetimes of the generated radical species changed as the solution pH changed. When the point of zero charge of the catalyst was higher than the solution pH, the surface of the catalyst displayed a positive charge; thus, it could adsorb more SO4•− [41]. In addition, the lifetimes of HO and SO4•− decreased under alkaline conditions; thus, the radical species that diffused into the bulk phase might have been insufficient for further degradation [42].

3.5. Applicability of HWWC

The stability of the catalyst is an important index for practical applications in wastewater treatment. Therefore, sequential OTC degradation experiments were performed to test the reusability of the catalyst (Figure 8a). The reaction was conducted for 24 min, after which the HWWC was recovered using an external magnet. The degradation performance of the HWWC did not significantly decrease after eight repeat experiments with a final degradation efficiency of 92.7%. This result indicated that the catalyst exhibited excellent regeneration performance and stability. Since the actual environmental water contains a large amount of organic and inorganic compounds, the activity of SO4•− and HO may be reduced. Therefore, it is necessary to investigate the effect of radical scavengers of these organic and inorganic compounds. Chloride ion (Cl), one of the representative inorganic compounds present in large amounts in environmental water, can reduce degradation efficiency by the following equations (Equations (10)–(11)) [38,43]. In the present condition, however, the effect of chloride ion ([Cl]0 = 10 mg/L) was negligible (Figure 8b). By contrast, the OTC degradation was significantly reduced by the organic compounds present in the secondary effluent (pH = 7.2, [DOC]0 = 4.71 mg/L, UV254 = 0.100, SUVA = 2.12) (Figure 8c). This is because the electron-rich moieties in the molecular structure of natural organic matter (NOM) present in the secondary effluent can be readily attacked by electrophilic radicals such as SO4•− and HO [44].
SO4•− + Cl → SO42− + Cl
Cl + HO → Cl + HO

4. Conclusions

HWWC was successfully prepared by a simple magnetic separation method. The XRD and SEM-EDS results revealed that the HWWC consisted of a mixture of γ-Fe2O3 and α-Fe2O3. The magnetic saturation of HWWC was sufficient to be separated by conventional magnets, which can facilitate their application for water treatment. The control experiment showed that OTC was removed by the generated radical species when both HWWC and PS were present. SO4•− was the dominant oxidizing species in the OTC degradation by persulfate activation using HWWC catalyst. Influencing parameters such as HWWC dose, PS concentration, and solution pH were evaluated, and the degradation efficiency of OTC increased with increasing HWWC dose and PS concentration, and the optimal pH values for OTC degradation were 3 and 4. In addition, the HWWC degraded OTC after eight repeat experiments with great stability. Degradation efficiency was significantly affected by NOM present in the secondary effluent, while the effect of chloride ions was negligible. Overall, these results suggest that PS activation using magnetic Fe2O3 catalysts derived from hand-warmer waste could be an effective alternative for removing OTC and other recalcitrant organic compounds in water.

Author Contributions

Conceptualization, Y.-J.L. and C.-G.L.; methodology, Y.-J.L.; validation, Y.-J.L.; formal analysis, Y.-J.L.; writing—original draft preparation, Y.-J.L.; writing—review and editing, C.-G.L., S.-J.P. and E.H.J.; visualization, Y.-J.L.; supervision, C.-G.L., S.-J.P. and E.H.J.; funding acquisition, E.H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was conducted with the support of the Cooperative Research Program for Agriculture Science and Technology Development (Project No. PJ01571602), Rural Development Administration, Korea.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, D.; Li, M.; Li, X.; Ren, F.; Sun, P.; Zhou, L. Core-shell Zn/Co MOFs derived Co3O4/CNTs as an efficient magnetic heterogeneous catalyst for persulfate activation and oxytetracycline degradation. Chem. Eng. J. 2020, 387, 124008. [Google Scholar] [CrossRef]
  2. Baquero, F.; Martinez, J.L.; Canton, R. Antibiotics and antibiotic resistance in water environments. Curr. Opin Biotechnol. 2008, 19, 260–265. [Google Scholar] [CrossRef]
  3. Wang, L.; Yang, H.; Zhang, C.; Mo, Y.; Lu, X. Determination of oxytetracycline, tetracycline and chloramphenicol antibiotics in animal feeds using subcritical water extraction and high performance liquid chromatography. Anal. Chim. Acta 2008, 619, 54–58. [Google Scholar] [CrossRef]
  4. Uslu, M.O.; Balcioglu, I.A. Simultaneous removal of oxytetracycline and sulfamethazine antibacterials from animal waste by chemical oxidation processes. J. Agric. Food Chem. 2009, 57, 11284–11291. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, Y.; He, X.; Fu, Y.; Dionysiou, D.D. Kinetics and mechanism investigation on the destruction of oxytetracycline by UV-254nm activation of persulfate. J. Hazard. Mater. 2016, 305, 229–239. [Google Scholar] [CrossRef]
  6. Kemper, N. Veterinary antibiotics in the aquatic and terrestrial environment. Ecol. Indic. 2008, 8, 1–13. [Google Scholar] [CrossRef]
  7. Sarmah, A.K.; Meyer, M.T.; Boxall, A.B. A global perspective on the use, sales, exposure pathways, occurrence, fate and effects of veterinary antibiotics (VAs) in the environment. Chemosphere 2006, 65, 725–759. [Google Scholar] [CrossRef]
  8. Kummerer, K. Antibiotics in the aquatic environment—A review—Part I. Chemosphere 2009, 75, 417–434. [Google Scholar] [CrossRef]
  9. Zheng, S.; Qiu, X.; Chen, B.; Yu, X.; Liu, Z.; Zhong, G.; Li, H.; Chen, M.; Sun, G.; Huang, H.; et al. Antibiotics pollution in Jiulong River estuary: Source, distribution and bacterial resistance. Chemosphere 2011, 84, 1677–1685. [Google Scholar] [CrossRef]
  10. Li, B.; Zhang, T. Mass flows and removal of antibiotics in two municipal wastewater treatment plants. Chemosphere 2011, 83, 1284–1289. [Google Scholar] [CrossRef]
  11. Li, W.; Shi, Y.; Gao, L.; Liu, J.; Cai, Y. Occurrence and removal of antibiotics in a municipal wastewater reclamation plant in Beijing, China. Chemosphere 2013, 92, 435–444. [Google Scholar] [CrossRef]
  12. Xu, J.; Xu, Y.; Wang, H.; Guo, C.; Qiu, H.; He, Y.; Zhang, Y.; Li, X.; Meng, W. Occurrence of antibiotics and antibiotic resistance genes in a sewage treatment plant and its effluent-receiving river. Chemosphere 2015, 119, 1379–1385. [Google Scholar] [CrossRef]
  13. Golovko, O.; Kumar, V.; Fedorova, G.; Randak, T.; Grabic, R. Seasonal changes in antibiotics, antidepressants/psychiatric drugs, antihistamines and lipid regulators in a wastewater treatment plant. Chemosphere 2014, 111, 418–426. [Google Scholar] [CrossRef]
  14. Chen, J.; Hu, Y.; Huang, W.; Liu, Y.; Tang, M.; Zhang, L.; Sun, J. Biodegradation of oxytetracycline and electricity generation in microbial fuel cell with in situ dual graphene modified bioelectrode. Bioresour. Technol. 2018, 270, 482–488. [Google Scholar] [CrossRef]
  15. Garrido-Ramírez, E.G.; Theng, B.K.G.; Mora, M.L. Clays and oxide minerals as catalysts and nanocatalysts in Fenton-like reactions—A review. Appl. Clay Sci. 2010, 47, 182–192. [Google Scholar] [CrossRef]
  16. Feng, Y.; Wu, D.; Liao, C.; Deng, Y.; Zhang, T.; Shih, K. Red mud powders as low-cost and efficient catalysts for persulfate activation: Pathways and reusability of mineralizing sulfadiazine. Sep. Purif. Technol. 2016, 167, 136–145. [Google Scholar] [CrossRef]
  17. Liang, C.J.; Bruell, C.J.; Marley, M.C.; Sperry, K.L. Thermally Activated Persulfate Oxidation of Trichloroethylene (TCE) and 1,1,1-Trichloroethane (TCA) in Aqueous Systems and Soil Slurries. Soil Sediment Contam. Int. J. 2010, 12, 207–228. [Google Scholar] [CrossRef]
  18. Zhang, T.; Chen, Y.; Wang, Y.; Le Roux, J.; Yang, Y.; Croue, J.P. Efficient peroxydisulfate activation process not relying on sulfate radical generation for water pollutant degradation. Environ. Sci. Technol. 2014, 48, 5868–5875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Abu Amr, S.S.; Aziz, H.A.; Adlan, M.N. Optimization of stabilized leachate treatment using ozone/persulfate in the advanced oxidation process. Waste Manag. 2013, 33, 1434–1441. [Google Scholar] [CrossRef]
  20. Peng, L.; Deng, D.; Guan, M.; Fang, X.; Zhu, Q. Remediation HCHs POPs-contaminated soil by activated persulfate technologies: Feasibility, impact of activation methods and mechanistic implications. Sep. Purif. Technol. 2015, 150, 215–222. [Google Scholar] [CrossRef]
  21. Liu, C.S.; Shih, K.; Sun, C.X.; Wang, F. Oxidative degradation of propachlor by ferrous and copper ion activated persulfate. Sci. Total Environ. 2012, 416, 507–512. [Google Scholar] [CrossRef]
  22. Teel, A.L.; Ahmad, M.; Watts, R.J. Persulfate activation by naturally occurring trace minerals. J. Hazard. Mater. 2011, 196, 153–159. [Google Scholar] [CrossRef]
  23. Fang, G.-D.; Dionysiou, D.D.; Al-Abed, S.R.; Zhou, D.-M. Superoxide radical driving the activation of persulfate by magnetite nanoparticles: Implications for the degradation of PCBs. Appl. Catal. B: Environ. 2013, 129, 325–332. [Google Scholar] [CrossRef]
  24. Ioannidi, A.; Oulego, P.; Collado, S.; Petala, A.; Arniella, V.; Frontistis, Z.; Angelopoulos, G.N.; Diaz, M.; Mantzavinos, D. Persulfate activation by modified red mud for the oxidation of antibiotic sulfamethoxazole in water. J. Environ. Manag. 2020, 270, 110820. [Google Scholar] [CrossRef] [PubMed]
  25. Naim, S.; Ghauch, A. Ranitidine abatement in chemically activated persulfate systems: Assessment of industrial iron waste for sustainable applications. Chem. Eng. J. 2016, 288, 276–288. [Google Scholar] [CrossRef]
  26. Lian, J.-Z.; Tsai, C.-T.; Chang, S.-H.; Lin, N.-H.; Hsieh, Y.-H. Iron waste as an effective depend on TiO2 for photocatalytic degradation of dye waste water. Optik 2017, 140, 197–204. [Google Scholar] [CrossRef]
  27. Yu, Y.; Paul Chen, J. Key factors for optimum performance in phosphate removal from contaminated water by a Fe-Mg-La tri-metal composite sorbent. J. Colloid. Interface Sci. 2015, 445, 303–311. [Google Scholar] [CrossRef]
  28. Qiu, J.; Yang, R.; Li, M.; Jiang, N. Preparation and characterization of porous ultrafine Fe2O3 particles. Mater. Res. Bull. 2005, 40, 1968–1975. [Google Scholar] [CrossRef]
  29. Molodtsova, T.; Gorshenkov, M.; Saliev, A.; Vanyushin, V.; Goncharov, I.; Smirnova, N. One-step synthesis of γ-Fe2O3/Fe3O4 nanocomposite for sensitive electrochemical detection of hydrogen peroxide. Electrochim. Acta 2021, 370, 137723. [Google Scholar] [CrossRef]
  30. Khoshnam, M.; Salimijazi, H. Synthesis and characterization of magnetic-photocatalytic Fe3O4/SiO2/a-Fe2O3 nano core-shell. Surf. Interfaces 2021, 26, 101322. [Google Scholar] [CrossRef]
  31. Chu, J.-H.; Kang, J.-K.; Park, S.-J.; Lee, C.-G. Application of magnetic biochar derived from food waste in heterogeneous sono-Fenton-like process for removal of organic dyes from aqueous solution. J. Water Process Eng. 2020, 37, 101455. [Google Scholar] [CrossRef]
  32. Chu, J.-H.; Kang, J.-K.; Park, S.-J.; Lee, C.-G. Enhanced sonocatalytic degradation of bisphenol A with a magnetically recoverable biochar composite using rice husk and rice bran as substrate. J. Environ. Chem. Eng. 2021, 9, 105284. [Google Scholar] [CrossRef]
  33. Zouanti, M.; Bezzina, M.; Dhib, R. Experimental study of degradation and biodegradability of oxytetracycline antibiotic in aqueous solution using Fenton process. Environ. Eng. Res. 2019, 25, 316–323. [Google Scholar] [CrossRef] [Green Version]
  34. Ma, Q.; Zhang, X.; Guo, R.; Zhang, H.; Cheng, Q.; Xie, M.; Cheng, X. Persulfate activation by magnetic γ-Fe2O3/Mn3O4 nanocomposites for degradation of organic pollutants. Sep. Purif. Technol. 2019, 210, 335–342. [Google Scholar] [CrossRef]
  35. Guan, Y.H.; Ma, J.; Ren, Y.M.; Liu, Y.L.; Xiao, J.Y.; Lin, L.Q.; Zhang, C. Efficient degradation of atrazine by magnetic porous copper ferrite catalyzed peroxymonosulfate oxidation via the formation of hydroxyl and sulfate radicals. Water Res. 2013, 47, 5431–5438. [Google Scholar] [CrossRef]
  36. Pham, A.N.; Xing, G.; Miller, C.J.; Waite, T.D. Fenton-like copper redox chemistry revisited: Hydrogen peroxide and superoxide mediation of copper-catalyzed oxidant production. J. Catal. 2013, 301, 54–64. [Google Scholar] [CrossRef]
  37. Liang, H.-Y.; Zhang, Y.-Q.; Huang, S.-B.; Hussain, I. Oxidative degradation of p-chloroaniline by copper oxidate activated persulfate. Chem. Eng. J. 2013, 218, 384–391. [Google Scholar] [CrossRef]
  38. Norzaee, S.; Taghavi, M.; Djahed, B.; Kord Mostafapour, F. Degradation of Penicillin G by heat activated persulfate in aqueous solution. J. Environ. Manag. 2018, 215, 316–323. [Google Scholar] [CrossRef]
  39. Liang, C.; Lee, I.L. In situ iron activated persulfate oxidative fluid sparging treatment of TCE contamination--a proof of concept study. J. Contam. Hydrol. 2008, 100, 91–100. [Google Scholar] [CrossRef]
  40. Chen, Y.; Yan, J.; Ouyang, D.; Qian, L.; Han, L.; Chen, M. Heterogeneously catalyzed persulfate by CuMgFe layered double oxide for the degradation of phenol. Appl. Catal. A: Gen. 2017, 538, 19–26. [Google Scholar] [CrossRef]
  41. Long, M.; Brame, J.; Qin, F.; Bao, J.; Li, Q.; Alvarez, P.J. Phosphate Changes Effect of Humic Acids on TiO2 Photocatalysis: From Inhibition to Mitigation of Electron-Hole Recombination. Environ. Sci. Technol. 2017, 51, 514–521. [Google Scholar] [CrossRef] [PubMed]
  42. Li, Z.; Chen, Z.; Xiang, Y.; Ling, L.; Fang, J.; Shang, C.; Dionysiou, D.D. Bromate formation in bromide-containing water through the cobalt-mediated activation of peroxymonosulfate. Water Res. 2015, 83, 132–140. [Google Scholar] [CrossRef] [PubMed]
  43. Chu, J.-H.; Kang, J.-K.; Park, S.-J.; Lee, C.-G. Ultrasound-activated peroxydisulfate process with copper film to remove bisphenol A: Operational parameter impact and back propagation-artificial neural network modeling. J. Water Process Eng. 2021, 44, 102326. [Google Scholar] [CrossRef]
  44. Ma, J.; Yang, Y.; Jiang, X.; Xie, Z.; Li, X.; Chen, C.; Chen, H. Impacts of inorganic anions and natural organic matter on thermally activated persulfate oxidation of BTEX in water. Chemosphere 2018, 190, 296–306. [Google Scholar] [CrossRef]
Figure 1. (a) pHpzc measurement and (b) X-ray diffraction pattern of prepared hand-warmer waste catalyst.
Figure 1. (a) pHpzc measurement and (b) X-ray diffraction pattern of prepared hand-warmer waste catalyst.
Applsci 11 10447 g001
Figure 2. (a) Scanning electron microscope image and (b) energy dispersive X-ray spectrometer result of the prepared hand-warmer waste catalyst.
Figure 2. (a) Scanning electron microscope image and (b) energy dispersive X-ray spectrometer result of the prepared hand-warmer waste catalyst.
Applsci 11 10447 g002
Figure 3. Vibrating sample magnetometer (VSM) analysis result of HWWC.
Figure 3. Vibrating sample magnetometer (VSM) analysis result of HWWC.
Applsci 11 10447 g003
Figure 4. Control experiment for oxytetracycline (OTC) degradation ([PS]0 = 1 mM, [OTC]0 = 20 μM, [HWWC]0 = 0.2 g/L, pH = 6). PS: persulfate; HWWC: hand-warmer waste catalyst.
Figure 4. Control experiment for oxytetracycline (OTC) degradation ([PS]0 = 1 mM, [OTC]0 = 20 μM, [HWWC]0 = 0.2 g/L, pH = 6). PS: persulfate; HWWC: hand-warmer waste catalyst.
Applsci 11 10447 g004
Figure 5. Oxytetracycline (OTC) degradation in the presence of scavengers ([OTC]0 = 20 μM, [PS]0 = 1 mM, [HWWC]0 = 0.2 g/L, [scavenger]0 = 10 mg/L, pH = 6). PS: persulfate; HWWC: hand-warmer waste catalyst.
Figure 5. Oxytetracycline (OTC) degradation in the presence of scavengers ([OTC]0 = 20 μM, [PS]0 = 1 mM, [HWWC]0 = 0.2 g/L, [scavenger]0 = 10 mg/L, pH = 6). PS: persulfate; HWWC: hand-warmer waste catalyst.
Applsci 11 10447 g005
Figure 6. Oxytetracycline (OTC) degradation with (a) different catalyst dosages and (b) different persulfate concentrations ([OTC]0 = 20 μM).
Figure 6. Oxytetracycline (OTC) degradation with (a) different catalyst dosages and (b) different persulfate concentrations ([OTC]0 = 20 μM).
Applsci 11 10447 g006
Figure 7. Oxytetracycline (OTC) degradation efficiency with different pH conditions ([PS]0 = 1 mM, [OTC]0 = 20 μM, [HWWC]0 = 0.2 g/L). PS: persulfate; HWWC: hand-warmer waste catalyst.
Figure 7. Oxytetracycline (OTC) degradation efficiency with different pH conditions ([PS]0 = 1 mM, [OTC]0 = 20 μM, [HWWC]0 = 0.2 g/L). PS: persulfate; HWWC: hand-warmer waste catalyst.
Applsci 11 10447 g007
Figure 8. (a) Sequential oxytetracycline (OTC) degradation test ([PS]0 = 1 mM, [OTC]0 = 20 μM, [HWWC]0 = 0.2 g/L, pH = 6); (b) OTC degradation in the presence of Cl ([Cl]0 = 10 mg/L); (c) Degradation kinetics of OTC by persulfate activation using HWWC in the secondary effluent (pH = 7.2, [DOC]0 = 4.71 mg/L, UV254 = 0.100, SUVA = 2.12). PS: persulfate; HWWC: hand-warmer waste catalyst.
Figure 8. (a) Sequential oxytetracycline (OTC) degradation test ([PS]0 = 1 mM, [OTC]0 = 20 μM, [HWWC]0 = 0.2 g/L, pH = 6); (b) OTC degradation in the presence of Cl ([Cl]0 = 10 mg/L); (c) Degradation kinetics of OTC by persulfate activation using HWWC in the secondary effluent (pH = 7.2, [DOC]0 = 4.71 mg/L, UV254 = 0.100, SUVA = 2.12). PS: persulfate; HWWC: hand-warmer waste catalyst.
Applsci 11 10447 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lee, Y.-J.; Lee, C.-G.; Park, S.-J.; Jho, E.H. Degradation of Oxytetracycline by Persulfate Activation Using a Magnetic Separable Iron Oxide Catalyst Derived from Hand-Warmer Waste. Appl. Sci. 2021, 11, 10447. https://doi.org/10.3390/app112110447

AMA Style

Lee Y-J, Lee C-G, Park S-J, Jho EH. Degradation of Oxytetracycline by Persulfate Activation Using a Magnetic Separable Iron Oxide Catalyst Derived from Hand-Warmer Waste. Applied Sciences. 2021; 11(21):10447. https://doi.org/10.3390/app112110447

Chicago/Turabian Style

Lee, Youn-Jun, Chang-Gu Lee, Seong-Jik Park, and Eun Hea Jho. 2021. "Degradation of Oxytetracycline by Persulfate Activation Using a Magnetic Separable Iron Oxide Catalyst Derived from Hand-Warmer Waste" Applied Sciences 11, no. 21: 10447. https://doi.org/10.3390/app112110447

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