Next Article in Journal
Investigation on Adsorption and Decomposition Properties of CL-20/FOX-7 Molecules on MgH2(110) Surface by First-Principles
Next Article in Special Issue
Solubility Data of the Bioactive Compound Piperine in (Transcutol + Water) Mixtures: Computational Modeling, Hansen Solubility Parameters and Mixing Thermodynamic Parameters
Previous Article in Journal
Erratum: Moreira, J., et al., Spin-Coated Polysaccharide-Based Multilayered Freestanding Films with Adhesive and Bioactive Moieties. Molecules 2020, 25, 840
Previous Article in Special Issue
Solubility Data and Computational Modeling of Baricitinib in Various (DMSO + Water) Mixtures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Activation of Peracetic Acid with Lanthanum Cobaltite Perovskite for Sulfamethoxazole Degradation under a Neutral pH: The Contribution of Organic Radicals

State Key Laboratory of Pollution Control and Resources Reuse, College of Environmental Science and Engineering, Tongji University, Shanghai 200092, China
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(12), 2725; https://doi.org/10.3390/molecules25122725
Submission received: 25 May 2020 / Revised: 9 June 2020 / Accepted: 11 June 2020 / Published: 12 June 2020
(This article belongs to the Special Issue 25th Anniversary of Molecules—Recent Advances in Physical Chemistry)

Abstract

:
Advanced oxidation processes (AOPs) are effective ways to degrade refractory organic contaminants, relying on the generation of inorganic radicals (e.g., OH and SO4•−). Herein, a novel AOP with organic radicals (R-O) was reported to degrade contaminants. Lanthanum cobaltite perovskite (LaCoO3) was used to activate peracetic acid (PAA) for organic radical generation to degrade sulfamethoxazole (SMX). The results show that LaCoO3 exhibited an excellent performance on PAA activation and SMX degradation at neutral pH, with low cobalt leaching. Meanwhile, LaCoO3 also showed an excellent reusability during PAA activation. In-depth investigation confirmed CH3C(O)O and CH3C(O)OO as the key reactive species for SMX degradation in LaCoO3/PAA system. The presence of Cl (1–100 mM) slightly inhibited the degradation of SMX in the LaCoO3/PAA system, whereas the addition of HCO3 (0.1–1 mM) and humic aid (1–10 mg/L) could significantly inhibit SMX degradation. This work highlights the generation of organic radicals via the heterogeneous activation of PAA and thus provides a promising way to destruct contaminants in wastewater treatment.

1. Introduction

Advanced oxidation processes (AOPs) have been extensively used to degrade toxic and bio-recalcitrant organic contaminants in wastewater treatment [1,2,3,4,5,6,7,8]. The commonly used AOPs are generally based on the in situ generation of highly reactive inorganic radicals, such as the hydroxyl radical (HO) and the sulfate radical (SO4•−). They are usually generated through the homogeneous cleavage of the peroxide bond in inorganic peroxides, such as hydrogen peroxide (H2O2), peroxydisulfate (PDS) and peroxymonosulfate (PMS) [9]. The inorganic radicals possess high oxidation potential (E0 = 2.5–3.1 V for SO4•- and E0 = 1.9–2.7 V for HO) [10] and thus can effectively destruct a wide range of organic contaminants with the second-order rate constants ranging from 105 to 109 M−1 s−1 for SO4•− [11] and from 107 to 1010 M−1 s−1 for HO [12]. However, the short half-life (30–40 μs for SO4•−, 1 μs for HO) of inorganic radicals may reduce the possibility of their interaction with the target contaminants [13,14]. On the other hand, the non-selectivity of inorganic radicals render them susceptible to consumption by the water matrices, e.g., anions and natural organic matters. These drawbacks reduce the degradation efficiency of target contaminants by inorganic radicals in real wastewater treatment [15].
Recently, an emerging AOP mainly based on the organic radicals (R-O) has become an attractive alternative for organic contaminant degradation in wastewater treatment. Organic radicals are commonly generated from the activation of peroxide bond in the organic peroxy acid, such as peracetic acid (PAA, CH3C(O)OOH) [16,17]. PAA is known as a broad-spectrum antimicrobial agent and has been widely used as an alternative to conventional chlorine disinfectants in wastewater disinfection, owing to the advantages of PAA, such as strong oxidation power (redox potential ranging from 1.06 to 1.96 V) [18] and a low potential of harmful disinfection byproduct generation [19,20,21]. PAA is produced by the reaction of acetic acid and H2O2 in the presence of a strong acid catalyst, such as sulfuric acid [22,23,24], and thus the commercial PAA solution is an equilibrium solution with acetic acid and H2O2 according to Equation (1).
CH 3 C ( = O ) OOH + H 2 O CH 3 C ( = O ) OH + H 2 O 2
PAA can directly oxidize some organic contaminants in wastewater treatment, e.g., amino acid [25] and β-lactam antibiotics [26]. After the introduction of intensive energy (e.g., heat and UV) or catalysts, the peroxide bond in PAA can be activated to produce reactive species (e.g., HO and R-O) [19,27,28,29]. Compared with the inorganic peroxides, e.g., H2O2 (213 kJ mol−1), PAA has a weaker -O–O- bond (159 kJ mol−1 for PAA [26,30]), which can be easily activated for intensive radical generation to attack target contaminants, such as pharmaceuticals [19,31], phenols [27,32], and dyes [33]. Therefore, PAA shows a great potential to be an ideal alternative to H2O2 in AOPs and the development of efficient and environmentally friendly methods to activate PAA is of great significance.
Many attempts have been made to develop effective strategies for the activation of PAA. Among them, transition metal ions have been reported as the most viable activators. For example, iron and cobalt ions showed a great performance for the activation of PAA to degrade organic contaminants [31,34]. However, one of the serious hindrances to this homogeneous processes was the poor reusability and potential toxicity of some metal ions [35]. A heterogeneous catalyst containing cobalt with low cobalt leaching and high reusability is expected to overcome these problems. In this work, a Co-based perovskite (LaCoO3) was used to activate PAA to eliminate recalcitrant contaminants. Perovskite, with the ABO3 structure, has recently attracted considerable interests in the field of material science and heterogeneous catalysis [36,37], due to the flexible chemical composition, element abundance and structural stability. In particular, perovskite oxides are regarded as excellent catalysts in wastewater treatment. For example, LaCoO3 has been used as a catalyst to activate PMS or H2O2 for contaminant degradation [36,38,39,40]. However, the performance of LaCoO3 on PAA activation and the mechanisms underlying the LaCoO3activated PAA reaction are still unknown.
Herein, LaCoO3 perovskite was synthesized via the sol-gel method and then used to activate PAA to degrade sulfamethoxazole (SMX). The aims of this study were: 1) to investigate the performance of LaCoO3/PAA on the degradation of SMX; 2) to evaluate the impact of water matrices on SMX degradation in the activated PAA process; 3) to identify the reactive radical species dominated in the contaminant degradation; 4) to reveal the activation mechanism of PAA by LaCoO3. To our best knowledge, this study is among the first to investigate the activation of PAA by perovskite and elucidate the mechanism for organic radical generation.

2. Results

2.1. Characterization of LaCoO3

Figure 1a shows the X-ray powder diffraction (XRD) pattern of LaCoO3. The diffraction peaks at 2θ values of 23.30°, 32.83°, 33.29°, 39.76°, 40.13°, 47.68°, 52.71°, 53.66°, 59.08°, 59.93° and 68.95° correspond to miller indices values of (012), (110), (104), (202), (006), (024), (122), (116), (214), (018) and (220), respectively [38]. In addition, the morphological properties of LaCoO3 was analyzed by Transmission Electron Microscope (TEM). The TEM image (Figure 1b) shows a disordered microstructure of the porous LaCoO3 nanospheres, with the average particle size approximately 23 nm according to the transmission electron microscope (TEM) image. The selected-area electron diffraction (SAED) pattern in Figure 1c revealed that the LaCoO3 nanospheres were polycrystalline. The Brunauer–Emmett–Teller (BET) area of the sample was obtained to be 14 m2/g and the zeta potential of LaCoO3 was determined to be 6.1.

2.2. Degradation of SMX by PAA Activated with LaCoO3

The prepared LaCoO3 was used to activate PAA for the degradation of SMX. As shown in Figure 2, the loss of SMX was not observed in the presence of LaCoO3 alone, indicating the negligible adsorption of SMX on LaCoO3. SMX was slightly degraded after the addition of PAA alone, suggesting the slight reactivity of PAA towards SMX. However, SMX could be almost completely degraded within 60 min in the presence of PAA and LaCoO3; hence, LaCoO3 and PAA showed a synergistic effect on SMX degradation. It should be addressed that the removal efficiency of SMX achieved by LaCoO3/PAA in this study was comparable or even higher than many published studies [41,42,43,44,45,46], and with much less catalyst addition. For example, Yan et al. [46] prepared CuO@Al2O3 to activate PMS and approximately 90% of SMX was removed under similar conditions ([SMX]0 = 39.5 μM, [PMS]0 = 0.4 mM, pH 6.2) but a large catalyst dosage ([cata]0 = 500 mg/L). Lalas et al. [45] reported that complete SMX degradation was obtained within 90 min through persulfate with 2 g immobilized CuOx catalyst. The LaCoO3 was reported as an excellent catalyst for the activation of peroxides, such as H2O2 and PMS, to generate reactive species to degrade contaminants [39,47,48]. It was most likely to activate PAA to degrade SMX. It is noted that commercial PAA solution always contains a certain amount of H2O2 during preparation; we thus evaluated the contribution of H2O2 to SMX degradation in the LaCoO3/PAA system. As Figure 2 shows, the degradation of SMX was negligible in the presence of H2O2, or H2O2/LaCoO3, where H2O2 concentration was equivalent to that in PAA solution. Hence, the degradation of SMX in LaCoO3/PAA was not affected by H2O2 in PAA solution. Meanwhile, this result also indicated that PAA was more easily activated by LaCoO3 than H2O2. Indeed, the peroxide bond energy in PAA was reported to be 159 kJ mol−1, which is lower than that in H2O2 (213 kJ mol−1) [31]. Hence, the dissociation of peroxide bond in PAA to generate reactive species is thermally feasible and even likely more easily than H2O2.

2.3. Effects of Catalyst Loadings

The effect of catalyst loading on the SMX degradation in the LaCoO3/PAA system was evaluated and the results are shown in Figure 3. Noting that the adsorption of SMX on LaCoO3 was negligible even in the presence of 0.1g/L LaCoO3 (data not shown), the removal of SMX could be attributed to the oxidative degradation rather than adsorption in the LaCoO3/PAA system. The degradation of SMX was gradually increased with the increasing loadings of LaCoO3 from 10 to 50 mg/L. A higher loading of LaCoO3 could supply more active sites for PAA to generate reactive species for contaminant degradation. When the LaCoO3 loading further increased to 100 mg/L, the degradation of SMX was only slightly increased. Hence, high loading of LaCoO3 had excess reactive sites for PAA activation, and even likely to quench the reactive species generated from PAA activation. Therefore, the excess loading of LaCoO3 could not further enhance the PAA activation.

2.4. Activation Mechanism

To get insight into the degradation of SMX in the LaCoO3/PAA system, the reactive species generated from PAA activation by LaCoO3 was comprehensively studied. Ascorbic acid (AA) is a common scavenger for reactive radicals, and thus is widely used for investigating radical contributions in contaminant degradation [33]. As can be seen in Figure 4a, the presence of AA significantly inhibited the degradation of SMX in the LaCoO3/PAA system, and the degradation rates decreased from 98.9% to 3.8% with the concentration of AA increased from 0 to 0.5 mM. This result indicated that the radical species contributed to the degradatoin of SMX in the LaCoO3/PAA system.
Generally, the activation of PAA could generate various reactive radical species, mainly including OH, CH3C(=O)O, CH3C(=O)OO, CH3 and CH3O2 (Equations (2–6)) [30].
CH 3 C ( = O ) OOH     CH 3 C ( = O ) O + OH
CH 3 C ( = O ) O     CH 3 + CO 2
CH 3 + O 2     CH 3 O 2
CH 3 C ( = O ) OOH + OH     CH 3 C ( = O ) OO + H 2 O
CH 3 C ( = O ) OOH + CH 3 C ( = O ) O     CH 3 C ( = O ) OO + CH 3 COOH
Some radical species might contribute to the contaminant degradation in the activated PAA system. For example, OH contributed to the degradation of carbamazepine and ibuprofen in the UV/PAA system, whereas a strong reactivity of CH3C(=O)O and/or CH3C(=O)OO was observed in the oxidation of certain naphthyl compounds, such as naproxen and 2-naphthoxyacetic acid [19].
The quenching experiments were further conducted to reveal which radicals were the main reactive species contributed to SMX degradation in the LaCoO3/PAA system. Tert-butanol (TBA) is a desired quenching agent to distinguish the contribution of OH in the activation of PAA because TBA is highly reactive towards OH (k = 6.0 × 108 M−1 s−1), but was inert towards other R-O. Noting that the rate constants between SMX and OH is 7.9 × 109 M−1 s−1, the contribution of OH can be almost completely consumed when the concentration of TBA is 1000 times higher than that of SMX. As shown in Figure 4a, the presence of excess TBA (500 mM) slightly inhibited the degradation of SMX in PAA/ LaCoO3 system, indicating that OH played a negligible role in LaCoO3/PAA system for SMX degradation. Moreover, MeOH is regarded as a common scavenger for OH (k = 6.0 × 108 M−1s−1), and has been recently reported as a scavenger for organic radicals (R-O) in the activated PAA [31]. Therefore, the addition of excess MeOH could be used to evaluate the contributions of R-O under the premise of TBA quenching results. In contrast to TBA, the addition of MeOH could significantly suppress the degradation of SMX in the LaCoO3/PAA system, indicating that the formed R-O might play a major role in SMX degradation.
To further validate the contributions of CH3 and CH3O2, N2 was pumped into the reaction solution. Because CH3 can quickly react with oxygen to form CH3O2 (Equation (4), k = (2.8–4.1) × 109 M−1 s−1), the contribution of CH3 and CH3O2 could be determined when the dissolved oxygen (DO) was excluded in the presence of high purity N2. The concentration of DO was determined to decrease from 4.95 mg/L to 0.21 mg/L after N2 purging. As shown in Figure 4b, the purging of N2 showed a negligible inhibition effect on SMX degradation, suggesting that the contributions of CH3 and CH3O2 to SMX degradation could be ignored. Therefore, CH3C(=O)O and CH3C(=O)OO were speculated to be the primary organic radical species that contributed to SMX degradation in the LaCoO3/PAA system.
The leaching of Co from LaCoO3 during the reaction was detected. After 60 min, the leached Co in the solution was determined to be 0.13 mg/L, which is much lower than the required concentration (< 1 mg/L) in the Environmental Quality Standards for Surface Water in China. In order to further investigate whether the leached Co contributed to the PAA activation for SMX degradation, PAA was introduced to the leached solution, and the result show that the degradation of SMX was not observed. Hence, the degradation of SMX in LaCoO3/PAA originated from the heterogeneous activation of PAA by LaCoO3 rather than the homogeneous activation with the leached Co ions. The heterogeneous activation of PAA with LaCoO3 was proposed in Figure 5. PAA was first adsorbed on the surface of LaCoO3, and then activated by the Co on the B site of perovskite to generate organic radicals, e.g., CH3COO, which could act as an oxidant for SMX degradation. Meanwhile, Co(II) was oxidized to Co(III) after PAA activation (Equation (7)) [31,49,50]. The decomposition of CH3COO can generate CO2 and CH3, which could further react with oxygen to generate CH3OO. The generated CH3 and CH3OO showed low reactivity towards contaminants. On the other hand, Co(III) further reacted with PAA to generate CH3COOO (Equation (8)), another effective reactive organic radical contributed to the rapid oxidation of SMX; while Co(III) was reduced to Co(II), and further participated in the PAA activation. Overall, Co acted as a catalyst role in the PAA activation to generate CH3C(=O)O and CH3C(=O)OO for SMX oxidation.
Co 2 + + CH 3 C ( = O ) OOH     Co 3 + + CH 3 C ( = O ) O + OH
Co 3 + + CH 3 C ( = O ) OOH     Co 2 + + CH 3 C ( = O ) OO + H +

2.5. Effects of Water Matrices

The effects of water matrices, such as common anions and natural organic matter, were evaluated on SMX degradation in the LaCoO3/PAA system. As shown in Figure 6a, the presence of Cl (0–20 mM), a pervasive anion in natural water, slightly inhibited the degradation of SMX in the LaCoO3/PAA system. Cl is a common radical scavenger in AOPs and can react with the radical species to generate less reactive chlorine radicals, such as Cl and Cl2•−. The consumption of reactive radicals by Cl could reduce the reaction between R-O with SMX and thus reduce the degradation efficiency of SMX.
HCO3/CO32− was reported as a common radical scavenger in PMS-based AOPs and the Co(II)/PAA system [51]; thus, the effects of HCO3/CO32− on the degradation of SMX was also studied in the LaCoO3/PAA system. As shown in Figure 6b, HCO3/CO32− could significantly inhibit the SMX oxidation in the LaCoO3/PAA system, and the inhibitory effect rapidly increased with an increasing concentration of HCO3/CO32−. Noting that the reaction was maintained at pH 7.0 with phosphate buffer, HCO3 is the dominant species in the system due to the pKa1 = 6.4. Although HCO3 can significantly quench OH, OH just played a minor role in the LaCoO3/PAA system for SMX degradation. Meanwhile, HCO3 was reported to show very low reactivity towards R-O, according to the phenomenon that the contaminant degradation in the UV/PAA + TBA system was not affected by HCO3 [19]. Thus, the reactions between HCO3 and the reactive radicals (OH and R-O) would not be the primary reason for the significant inhibition of SMX degradation in the LaCoO3/PAA system. The inhibitory effect could be attributed to the influence of HCO3 on the interaction between PAA and LaCoO3 by competitive adsorption.
The effects of natural organic matter on SMX degradation were studied by adding a certain amount of humic acid (HA) in the LaCoO3/PAA system. As shown in Figure 6c, the presence of HA (0–10 mg/L) showed an apparent inhibition on SMX degradation in the LaCoO3/PAA system. To be specific, the degradation of SMX significantly decreased from 100% to 51% after 50 min as the concentration of HA increased from 0 to 10 mg/L. HA was previously reported to react with R-O, with the rate constants reaching 104 L mg−1 s−1. Hence, the addition of HA could compete with SMX towards R-O and thus reduce the degradation efficiency of SMX in the LaCoO3/PAA system.
The impact of real water matrices on the degradation of SMX was further explored in LaCoO3/PAA. The characteristics of the surface water (SW) and wastewater (WW) were listed in Table 1. As shown in Figure 6d, the degradation of SMX was inhibited in SW and WW, with the degradation efficiency moderately decreased to 90% and 85% after 60 min, respectively. Hence, some matrices in the real water sample might impose an adverse effect on SMX degradation in the LaCoO3/PAA system. Based on the above analysis, the slight inhibition that occurred in SW and WW matrices was likely to primarily contribute to the scavenging of R-O by organic matters.

2.6. Reusability of LaCoO3

The reusability is crucial for the application of a catalyst in the heterogeneous reactions. LaCoO3 was reused to evaluate its performance on PAA activation after the reaction. As shown in Figure 7, an excellent performance of LaCoO3 in PAA activation to degrade SMX was still observed after four runs. Almost all SMX was completely removed in each run with a slight downward trend in the removal rates (from 98% in the 1st run to 95% in the 4th run). Moreover, the low leaching of Co under a neutral pH condition after the reaction by about 60 min also indicated the high structural stability of the prepared LaCoO3 during the reaction. Hence, the prepared LaCoO3 showed a high structural and chemical stability in PAA activation for the degradation of SMX.

3. Materials and Methods

3.1. Chemicals and Reagents

PAA solution (39% PAA and 6% H2O2 w/w), hydrogen peroxide solution (30% H2O2 w/w), sulfamethoxazole (SMX), cobaltous sulfate (CoSO4), cobaltous acetylacetonate (Co(C5H7O2)2), and cobaltic acetylacetonate (Co(C5H7O2)3) were purchased from Sigma-Aldrich (St. Louis, MO, USA). Oxone (KHSO5·0.5KHSO4·0.5K2SO4) and lanthanum nitrate hexahydrate (La(NO3)3·6H2O) were purchased from Aladdin Industrial Corporation. Cobalt nitrate hexahydrate (Co(NO3)2·6H2O), citric acid monohydrate (C6H8O7·H2O), sodium sulfate (Na2SO4), sodium nitrate (NaNO3), sodium chloride (NaCl), sodium bicarbonate (NaHCO3), ascorbic acid (AA), methanol (MeOH), and tert-butanol (TBA) were purchased from Sinopharm Chemical Reagent Co., Ltd., China. All solutions were prepared with ultrapure water from a Milli-Q academic (Millipore) system (MILLI-Q® INTEGRAL 5, 18.2 MΩ cm).

3.2. Preparation and Characterization of LaCoO3

LaCoO3 was synthesized by a sol-gel process. Briefly, the mixtures of metal nitrates and citric acid were stirred for 12 h and then heated at 100 °C to remove water in the oil bath. The resultant sticky gel was subsequently oven-dried at 100 °C overnight. The obtained spongy material was then grinded into fine powders and calcined for 7 h in a muffle furnace to form the desired perovskite structure. After cooling, the sample was washed and dried overnight.
Transmission electron microscope (TEM) and selected-area electron diffraction (SAED) patterns were conducted on the field emission transmission electron microscope (Tecnai G2 F30, Eindhoven, Holland). The X-ray powder diffraction (XRD) analysis was characterized by Cu Kα irradiation (λ = 1.5406 Å) (Bruker D8-Advance, Karlsruhe, Germany). The surface area of LaCoO3 was analyzed by a Brunauer–Emmett–Teller (BET) analyzer (Micrometrics ASAP 2020, Norcross, GA, USA). The zeta potential analysis was acquired through a Malvern Zetasizer with a Zetasizer (Nano ZS90, Malvern, UK) at 25 °C.

3.3. Experimental Procedures

The batch experiments were conducted in 100 mL amber glass bottles. The solution was mixed by magnetic stirring at room temperature (25 °C). The reactions were initiated by adding PAA (660 μM) to the solution containing SMX (50 μM) and LaCoO3 (20 mg/L). Sample aliquots were taken at the predetermined time intervals and subsequently quenched with excess Na2S2O3 immediately. Afterwards, the quenched sample was filtered through a 0.22 μm membrane and analyzed within 24 h. The initial pH of the solution was adjusted by H2SO4 and NaOH to the designated value. Control experiments without PAA or LaCoO3 were conducted to evaluate the contributions of PAA or LaCoO3 alone for the degradation of SMX. Radical scavenging experiments were conducted with addition of 500 mM of alcohols (tert-butyl alcohol (TBA) or methanol (MeOH)). In order to assess the reusability of LaCoO3 in the activation of PAA, the catalyst was collected by centrifugation and then washed with methanol and ultrapure water. Afterwards, the solid was dried at 60 °C overnight and four runs were carried out to evaluate the reusability. All the experiments were conducted in duplicate or more.

3.4. Analytical Methods

SMX was analyzed by high-performance liquid chromatography (HPLC 1290, Agilent Technology, Santa Clara, CA, USA) equipped with a Zorbax SB-C18 column (4.6 × 250 mm, 5 μm) and a UV detector at a wavelength of 265 nm. The flow rate was set at 1 mL min−1 and the injection was set at 20 μL. The mobile phase was the mixture of acetonitrile and water containing 0.4% acetic acid (70:30).
The PAA concentration in the solution was quantified by the N,N-diethyl-p-phenylenediamine (DPD) colorimetric method [19]. Briefly, the PAA solution was sampled and then was immediately mixed with excess KI (60 mM), DPD (2.8 mM), and phosphate buffer (0.5 M at pH 6.5) and measured at 515 nm on a UV-visible spectrophotometer (Beckman DU 520, Beckman Coulter, Inc., Fullerton, CA, USA).

4. Conclusions

LaCoO3 was successfully synthesized and used for PAA activation to degrade SMX. The rapid degradation of SMX was observed with the heterogeneous activation of PAA by LaCoO3 under a neutral pH condition. Cl was slightly inhibited by the degradation of SMX, whereas HCO3 and HA significantly reduce SMX in the LaCoO3/PAA system. LaCoO3 showed high structural stability and reusability after consecutive runs with low Co leaching under a neutral condition. Organic radicals, e.g., CH3C(=O)O and CH3C(=O)OO were proposed as the primary radical species responsible for SMX degradation in the LaCoO3/PAA system. This work provides an emerging AOP relying on organic radicals to degrade organic contaminants in the wastewater treatment.

Author Contributions

X.Z. and J.C. conceived and designed the experiments; H.W., B.L., and X.S. performed the experiments; H.W. and L.Z. analyzed the data; X.Z. and J.C. contributed reagents/materials/analysis tools; J.C. and L.Z. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Shanghai Science and Technology Committee (19DZ1208400), and the National Natural Science Foundation of China (51878431).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Oturan, M.A.; Aaron, J.-J. Advanced Oxidation Processes in Water Wastewater Treatment Principles and Applications A Review. Crit. Rev. Environ. Sci. Tech. 2014, 44, 2577–2641. [Google Scholar] [CrossRef]
  2. Zhou, D.; Chen, L.; Li, J.; Wu, F. Transition metal catalyzed sulfite auto-oxidation systems for oxidative decontamination in waters: A state-of-the-art minireview. Chem. Eng. J. 2018, 346, 726–738. [Google Scholar] [CrossRef]
  3. Xie, P.; Ma, J.; Liu, W.; Zou, J.; Yue, S.; Li, X.; Wiesner, M.R.; Fang, J. Removal of 2-MIB and geosmin using UV/persulfate: Contributions of hydroxyl and sulfate radicals. Water Res. 2015, 69, 223–233. [Google Scholar] [CrossRef] [PubMed]
  4. Lei, X.; You, M.; Pan, F.; Liu, M.; Yang, P.; Xia, D.; Li, Q.; Wang, Y.; Fu, J. CuFe2O4@GO nanocomposite as an effective and recoverable catalyst of peroxymonosulfate activation for degradation of aqueous dye pollutants. Chin. Chem. Lett. 2019, 30, 2216–2220. [Google Scholar] [CrossRef]
  5. Ma, M.; Chen, L.; Zhao, J.; Liu, W.; Ji, H. Efficient activation of peroxymonosulfate by hollow cobalt hydroxide for degradation of ibuprofen and theoretical study. Chin. Chem. Lett. 2019, 30, 2191–2195. [Google Scholar] [CrossRef]
  6. Liu, T.; Yin, K.; Liu, C.; Luo, J.; Crittenden, J.; Zhang, W.; Luo, S.; He, Q.; Deng, Y.; Liu, H. The role of reactive oxygen species and carbonate radical in oxcarbazepine degradation via UV, UV/H2O2: Kinetics, mechanisms and toxicity evaluation. Water Res. 2018, 147, 204–213. [Google Scholar] [CrossRef]
  7. Glaze, W.H.; Kang, J.W.; Chapin, D.H. The Chemistry of Water Treatment Processes Involving Ozone, Hydrogen Peroxide and Ultraviolet Radiation. Ozone Sci. Eng. 1987, 9, 335–352. [Google Scholar] [CrossRef]
  8. Glaze, W.H.; Kang, J.W. Advanced oxidation processes. Description of a kinetic model for the oxidation of hazardous materials in aqueous media with ozone and hydrogen peroxide in a semibatch reactor. Ind. Eng. Chem. Res. 1989, 28, 1573–1580. [Google Scholar] [CrossRef]
  9. Duan, P.; Qi, Y.; Feng, S.; Peng, X.; Wang, W.; Yue, Y.; Shang, Y.; Li, Y.; Gao, B.; Xu, X. Enhanced degradation of clothianidin in peroxymonosulfate/catalyst system via core-shell FeMn @ N-C and phosphate surrounding. Appl. Catal. B Environ. 2020, 267, 118717. [Google Scholar] [CrossRef]
  10. Rehman, F.; Sayed, M.; Khan, J.A.; Shah, N.S.; Khan, H.M.; Dionysiou, D.D. Oxidative removal of brilliant green by UV/S2O82−, UV/HSO5 and UV/H2O2 processes in aqueous media: A comparative study. J. Hazard. Mater. 2018, 357, 506–514. [Google Scholar] [CrossRef]
  11. Oh, W.-D.; Dong, Z.; Lim, T.-T. Generation of sulfate radical through heterogeneous catalysis for organic contaminants removal: Current development, challenges and prospects. Appl. Catal. B Environ. 2016, 194, 169–201. [Google Scholar] [CrossRef]
  12. Haag, W.R.; Yao, C.C.D. Rate constants for reaction of hydroxyl radicals with several drinking water contaminants. Environ. Sci. Tech. 1992, 26, 1005–1013. [Google Scholar] [CrossRef]
  13. Chen, J.B.; Fang, C.; Xia, W.J.; Huang, T.Y.; Huang, C.H. Selective Transformation of β-Lactam Antibiotics by Peroxymonosulfate: Reaction Kinetics and Non-Radical Mechanism. Environ. Sci. 2018, 52, 1461–1470. [Google Scholar] [CrossRef] [PubMed]
  14. Cao, J.Y.; Song, L. Degradation of tetracycline by peroxymonosulfate activated with zero-valent iron: Performance, intermediates, toxicity and mechanism. Chem. Eng. J. 2019, 364, 45–56. [Google Scholar] [CrossRef]
  15. Tsitonaki, A.; Petri, B.; Crimi, M.; Mosbk, H.; Siegrist, R.L.; Bjerg, P.L. In Situ Chemical Oxidation of Contaminated Soil and Groundwater Using Persulfate: A Review. Crit. Rev. Environ. Sci. Tech. 2010, 40, 55–91. [Google Scholar] [CrossRef]
  16. Kitis, M. Disinfection of wastewater with peracetic acid: A review. Environ. Int. 2004, 30, 47–55. [Google Scholar] [CrossRef]
  17. Chen, S.; Cai, M.Q.; Liu, Y.; Zhang, L.; Feng, L. Effects of water matrices on the degradation of naproxen by reactive radicals in the UV/peracetic acid process. Water Res. 2019, 150, 153–161. [Google Scholar] [CrossRef]
  18. Zhang, C.; Brown, P.J.B.; Hu, Z. Thermodynamic properties of an emerging chemical disinfectant, peracetic acid. Sci. Total. Environ. 2018, 621, 948–959. [Google Scholar] [CrossRef]
  19. Cai, M.; Sun, P.; Zhang, L.; Huang, C.H. UV/Peracetic Acid for Degradation of Pharmaceuticals and Reactive Species Evaluation. Environ Sci Technol. 2017, 51, 14217–14224. [Google Scholar] [CrossRef]
  20. Lee, W.N.; Huang, C.H. Formation of disinfection byproducts in wash water and lettuce by washing with sodium hypochlorite and peracetic acid sanitizers. Food Chem. X. 2019, 1, 100003. [Google Scholar] [CrossRef]
  21. Chhetri, R.K.; Baun, A.; Andersen, H.R. Acute toxicity and risk evaluation of the CSO disinfectants performic acid, peracetic acid, chlorine dioxide and their by-products hydrogen peroxide and chlorite. Sci. Total. Environ. 2019, 677, 1–8. [Google Scholar] [CrossRef]
  22. Shah, A.D.; Liu, Z.Q.; Salhi, E.; Höfer, T.; Werschkun, B.; von Gunten, U. Formation of disinfection by-products during ballast water treatment with ozone, chlorine, and peracetic acid: Influence of water quality parameters. Environ. Sci. Water Res. Tech. 2015, 1, 465–480. [Google Scholar]
  23. Zhao, X.; Zhang, T.; Zhou, Y.; Liu, D. Preparation of peracetic acid from hydrogen peroxide. J. Mol. Catal. A Chem. 2007, 271, 246–252. [Google Scholar] [CrossRef]
  24. Sun, P.Z.; Zhang, T.; Mejia-Tickner, B.; Zhang, R.; Cai, M.; Huang, C.H. Rapid Disinfection by Peracetic Acid Combined with UV Irradiation. Environ. Sci. Tech. Lett. 2018, 5, 400–404. [Google Scholar] [CrossRef]
  25. Du, P.; Liu, W.; Cao, H.; Zhao, H.; Huang, C.H. Oxidation of amino acids by peracetic acid: Reaction kinetics, pathways and theoretical calculations. Water Res X. 2018, 1, 100002. [Google Scholar] [CrossRef]
  26. Zhang, K.; Zhou, X.; Du, P.; Zhang, T.; Cai, M.; Sun, P.; Huang, C.H. Oxidation of beta-lactam antibiotics by peracetic acid: Reaction kinetics, product and pathway evaluation. Water Res. 2017, 123, 153–161. [Google Scholar] [CrossRef]
  27. Ina, E.V.R.; Makarova, K.; Golovina, E.A.; As, H.V.; Virkutyte, J. Free Radical Reaction Pathway, Thermochemistry of Peracetic Acid Homolysis, and Its Application for Phenol Degradation: Spectroscopic Study and Quantum Chemistry Calculations. Environ. Sci. Tech. 2010, 44, 6815–6821. [Google Scholar]
  28. Rothbart, S.; Ember, E.E.; Van, E.R. Mechanistic studies on the oxidative degradation of Orange II by peracetic acid catalyzed by simple manganese(ii) salts. Tuning the lifetime of the catalyst. New J. Chem. 2012, 36, 732–748. [Google Scholar] [CrossRef]
  29. Rokhina, E.V.; Makarova, K.; Lahtinen, M.; Golovina, E.A. Ultrasound-assisted MnO2 catalyzed homolysis of peracetic acid for phenol degradation: The assessment of process chemistry and kinetics. Chem. Eng. J. 2013, 221, 476–486. [Google Scholar] [CrossRef]
  30. Luukkonen, T.; Pehkonen, S.O. Peracids in water treatment: A critical review. Crit. Rev. Environ. Sci. Tech. 2016, 47, 1–39. [Google Scholar] [CrossRef] [Green Version]
  31. Wang, Z.; Wang, J.; Xiong, B.; Bai, F.; Wang, S.; Wan, Y.; Zhang, L.; Xie, P.; Wiesner, M.R. Application of Cobalt/Peracetic Acid to Degrade Sulfamethoxazole at Neutral Condition: Efficiency and Mechanisms. Environ. Sci. Tech. 2020, 54, 464–475. [Google Scholar] [CrossRef] [PubMed]
  32. Sharma, S.; Mukhopadhyay, M.; Murthy, Z.V.P. Degradation of 4-Chlorophenol in Wastewater by Organic Oxidants. Ind.Eng.Chem.Res. 2010, 49, 3094–3098. [Google Scholar] [CrossRef]
  33. Zhou, F.; Lu, C.; Yao, Y.; Sun, L.; Gong, F.; Li, D.; Pei, K.; Lu, W.; Chen, W. Activated carbon fibers as an effective metal-free catalyst for peracetic acid activation: Implications for the removal of organic pollutants. Chem. Eng. J. 2015, 281, 953–960. [Google Scholar] [CrossRef]
  34. Kim, J.; Zhang, T.; Liu, W.; Du, P.; Dobson, J.T.; Huang, C.H. Advanced Oxidation Process with Peracetic Acid and Fe(II) for Contaminant Degradation. Environ. Sci. Technol. 2019, 53, 13312–13322. [Google Scholar] [CrossRef]
  35. Sun, H.Q.; Kwan, C.K.; Suvorova, A.; Ming, H.; Moses, A. Catalytic oxidation of organic pollutants on pristine and surface nitrogen-modified carbon nanotubes with sulfate radicals. Appl. Catal. B Environ. 2014, 154–155, 134–141. [Google Scholar] [CrossRef]
  36. Guo, H.C.; Zhou, X.F.; Zhang, Y.L.; Yao, Q.F.; Qian, Y.J.; Chu, H.Q.; Chen, J.B. Carbamazepine degradation by heterogeneous activation of peroxymonosulfate with lanthanum cobaltite perovskite: Performance, mechanism and toxicity. J. Environ. Sci. 2020, 91, 10–21. [Google Scholar] [CrossRef] [PubMed]
  37. Wang, Y.; Ren, J.; Wang, Y.; Zhang, F.; Liu, X.; Guo, Y.; Lu, G. Nanocasted Synthesis of Mesoporous LaCoO3 Perovskite with Extremely High Surface Area and Excellent Activity in Methane Combustion. J. Phys. Chem. C. 2008, 112, 15293–15298. [Google Scholar] [CrossRef]
  38. Dacquin, J.P.; Dujardin, C.; Granger, P. Surface reconstruction of supported Pd on LaCoO3: Consequences on the catalytic properties in the decomposition of N2O. J. Catal. 2008, 253, 37–49. [Google Scholar] [CrossRef]
  39. Pang, X.; Guo, Y.; Zhang, Y.; Xu, B.; Qi, F. LaCoO3 perovskite oxide activation of peroxymonosulfate for aqueous 2-phenyl-5-sulfobenzimidazole degradation: Effect of synthetic method and the reaction mechanism. Chem. Eng. J. 2016, 304, 897–907. [Google Scholar] [CrossRef]
  40. Taran, O.P.; Ayusheev, A.B.; Ogorodnikova, O.L.; Prosvirin, I.P.; Isupova, L.A.; Parmon, V.N. Perovskite-like catalysts LaBO3 (B = Cu, Fe, Mn, Co, Ni) for wet peroxide oxidation of phenol. Appl. Catal. B Environ. 2016, 180, 86–93. [Google Scholar] [CrossRef]
  41. Li, C.X.; Wu, J.E.; Peng, W.; Fang, Z.F.; Liu, J. Peroxymonosulfate activation for efficient sulfamethoxazole degradation by Fe3O4/β-FeOOH nanocomposites: Coexistence of radical and non-radical reactions. Chem. Eng. J. 2019, 365, 904–914. [Google Scholar] [CrossRef]
  42. Wang, S.Z.; Xu, L.J.; Wang, J.L. Nitrogen-doped graphene as peroxymonosulfate activator and electron transfer mediator for the enhanced degradation of sulfamethoxazole. Chem. Eng. J. 2019, 375, 122041. [Google Scholar] [CrossRef]
  43. Gkika, C.; Petala, A.; Frontistis, Z. Heterogeneous activation of persulfate by lanthanum strontium cobaltite for sulfamethoxazole degradation. Catal. Today. 2020. [Google Scholar] [CrossRef]
  44. Oh, W.D.; Chang, V.W.C.; Lim, T.T. A comprehensive performance evaluation of heterogeneous bi2fe4o9/peroxymonosulfate system for sulfamethoxazole degradation. Environ. Sci. Pollut. Res. 2017, 26, 1026–1035. [Google Scholar] [CrossRef] [PubMed]
  45. Lalas, K.; Petala, A.; Frontistis, Z. Sulfamethoxazole degradation by the CuOx/persulfate system. Catal. Today. 2020. [Google Scholar] [CrossRef]
  46. Yan, J.F.; Li, J.; Peng, J.L. Efficient degradation of sulfamethoxazole by the CuO@Al2O3 (EPC) coupled PMS system: Optimization, degradation pathways and toxicity evaluation. Chem. Eng. J. 2019, 359, 1097–1110. [Google Scholar] [CrossRef]
  47. Hammouda, H.; Kalliola, S.; Sillanpaa, S.; Safaei, M.; Zhao, Z. Degradation and mineralization of phenol in aqueous medium by heterogeneous monopersulfate activation on nanostructured cobalt based-perovskite catalysts ACoO(3) (A = La, Ba, Sr and Ce): Characterization, kinetics and mechanism study. Appl. Catal. B Environ. 2017, 215, 60–73. [Google Scholar] [CrossRef]
  48. Duan, X.G.; Su, C.; Miao, J.; Zhong, Y.; Shao, Z.; Wang, S.; Sun, H. Insights into perovskite-catalyzed peroxymonosulfate activation: Maneuverable cobalt sites for promoted evolution of sulfate radicals. Appl. Catal. B Environ. 2018, 220, 626–634. [Google Scholar] [CrossRef]
  49. Hill, R.T. Decomposition of peracetic acid catalyzed by cobalt(II) and vanadium(V). Can. J. Chem. 1998, 76, 1064–1069. [Google Scholar]
  50. Kim, J.; Du, P.; Liu, W.; Luo, C.; Zhao, H.; Huang, C.H. Cobalt/Peracetic Acid: Advanced Oxidation of Aromatic Organic Compounds by Acetylperoxyl Radicals. Environ. Sci. Technol. 2020, 54, 5268–5278. [Google Scholar] [CrossRef]
  51. Zhang, Y.; Liu, J.; Moores, A.; Ghoshal, S. Transformation of Fluorotelomer Sulfonate by Cobalt(II)-Activated Peroxymonosulfate. Environ. Sci. Technol. 2020, 54, 4631–4640. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the lanthanum cobaltite perovskite are available from the authors.
Figure 1. (a) X-ray powder diffraction (XRD) pattern, (b) transmission electron microscope (TEM) image and (c) selected-area electron diffraction (SAED) image of the LaCoO3.
Figure 1. (a) X-ray powder diffraction (XRD) pattern, (b) transmission electron microscope (TEM) image and (c) selected-area electron diffraction (SAED) image of the LaCoO3.
Molecules 25 02725 g001
Figure 2. Sulfamethoxazole (SMX) degradation in different systems. LaCoO3 = 20 mg/L, SMX = 50 μM, peracetic acid (PAA) (or H2O2) = 660 μM, pH 7, 25 °C.
Figure 2. Sulfamethoxazole (SMX) degradation in different systems. LaCoO3 = 20 mg/L, SMX = 50 μM, peracetic acid (PAA) (or H2O2) = 660 μM, pH 7, 25 °C.
Molecules 25 02725 g002
Figure 3. Effect of LaCoO3 loading on SMX degradation in the LaCoO3/PAA system. SMX = 50 μM, PAA = 660 μM, pH 7, 25 °C.
Figure 3. Effect of LaCoO3 loading on SMX degradation in the LaCoO3/PAA system. SMX = 50 μM, PAA = 660 μM, pH 7, 25 °C.
Molecules 25 02725 g003
Figure 4. (a) Effect of ascorbic acid (AA), tert-butanol (TBA) and MeOH, and (b) Dissolved oxygen (DO) on the degradation of SMX in the LaCoO3/PAA system. LaCoO3 = 20 mg/L, SMX = 50 μM, PAA = 660 μM, pH 7, 25 °C.
Figure 4. (a) Effect of ascorbic acid (AA), tert-butanol (TBA) and MeOH, and (b) Dissolved oxygen (DO) on the degradation of SMX in the LaCoO3/PAA system. LaCoO3 = 20 mg/L, SMX = 50 μM, PAA = 660 μM, pH 7, 25 °C.
Molecules 25 02725 g004
Figure 5. Proposed mechanism of PAA activated by LaCoO3.
Figure 5. Proposed mechanism of PAA activated by LaCoO3.
Molecules 25 02725 g005
Figure 6. Effect of (a) Cl-, (b) HCO3, (c) humic acid and (d) real water on the degradation of SMX by the LaCoO3/PAA system. LaCoO3 = 20 mg/L, SMX = 50 μM, PAA = 660 μM, pH 7.0, 25 °C.
Figure 6. Effect of (a) Cl-, (b) HCO3, (c) humic acid and (d) real water on the degradation of SMX by the LaCoO3/PAA system. LaCoO3 = 20 mg/L, SMX = 50 μM, PAA = 660 μM, pH 7.0, 25 °C.
Molecules 25 02725 g006aMolecules 25 02725 g006b
Figure 7. Reusability of LaCoO3. LaCoO3 = 20 mg/L, SMX = 50 μM, PAA = 660 μM, pH 7.0, 25 °C. black square: 1st run, red circle: 2nd run, blue up triangle: 3rd run, magenta down triangle: 4th run.
Figure 7. Reusability of LaCoO3. LaCoO3 = 20 mg/L, SMX = 50 μM, PAA = 660 μM, pH 7.0, 25 °C. black square: 1st run, red circle: 2nd run, blue up triangle: 3rd run, magenta down triangle: 4th run.
Molecules 25 02725 g007
Table 1. Characteristics of water samples.
Table 1. Characteristics of water samples.
SampleTN (mg/L)TP (mg/L)pHUV254TOC (mg/L)Cl (mg/L)
SW0.810.077.80.133.132.6
WW11.10.847.10.4110.692.3
Note: SW: surface water; WW: wastewater.

Share and Cite

MDPI and ACS Style

Zhou, X.; Wu, H.; Zhang, L.; Liang, B.; Sun, X.; Chen, J. Activation of Peracetic Acid with Lanthanum Cobaltite Perovskite for Sulfamethoxazole Degradation under a Neutral pH: The Contribution of Organic Radicals. Molecules 2020, 25, 2725. https://doi.org/10.3390/molecules25122725

AMA Style

Zhou X, Wu H, Zhang L, Liang B, Sun X, Chen J. Activation of Peracetic Acid with Lanthanum Cobaltite Perovskite for Sulfamethoxazole Degradation under a Neutral pH: The Contribution of Organic Radicals. Molecules. 2020; 25(12):2725. https://doi.org/10.3390/molecules25122725

Chicago/Turabian Style

Zhou, Xuefei, Haowei Wu, Longlong Zhang, Bowen Liang, Xiaoqi Sun, and Jiabin Chen. 2020. "Activation of Peracetic Acid with Lanthanum Cobaltite Perovskite for Sulfamethoxazole Degradation under a Neutral pH: The Contribution of Organic Radicals" Molecules 25, no. 12: 2725. https://doi.org/10.3390/molecules25122725

Article Metrics

Back to TopTop