Next Article in Journal
Weathered Coal-Immobilized Microbial Materials as a Highly Efficient Adsorbent for the Removal of Lead
Previous Article in Journal
Polymer Nanoparticles with 2-HP-β-Cyclodextrin for Enhanced Retention of Uptake into HCE-T Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Degradation of Aqueous Oxytetracycline by an O3/CaO2 System in the Presence of HCO3: Performance, Mechanism, Degradation Pathways, and Toxicity Evaluation

1
School of Energy and Materials, Shanghai Polytechnic University, Shanghai 201209, China
2
College of Biology and the Environment, Nanjing Forestry University, Nanjing 210037, China
3
Anhui Jiuwu Tianhong Environmental Protection Technology Co., Ltd., Hefei 230011, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(3), 659; https://doi.org/10.3390/molecules29030659
Submission received: 27 December 2023 / Revised: 20 January 2024 / Accepted: 23 January 2024 / Published: 31 January 2024

Abstract

:
This research constructed a novel O3/CaO2/ HCO 3 system to degrade antibiotic oxytetracycline (OTC) in water. The results indicated that CaO2 and HCO 3 addition could promote OTC degradation in an O3 system. There is an optimal dosage of CaO2 (0.05 g/L) and HCO 3 (2.25 mmol/L) that promotes OTC degradation. After 30 min of treatment, approximately 91.5% of the OTC molecules were eliminated in the O3/CaO2/ HCO 3 system. A higher O3 concentration, alkaline condition, and lower OTC concentration were conducive to OTC decomposition. Active substances including ·OH, 1O2, · O 2 , and · HCO 3 play certain roles in OTC degradation. The production of ·OH followed the order: O3/CaO2/ HCO 3 > O3/CaO2 > O3. Compared to the sole O3 system, TOC and COD were easier to remove in the O3/CaO2/ HCO 3 system. Based on DFT and LC-MS, active species dominant in the degradation pathways of OTC were proposed. Then, an evaluation of the toxic changes in intermediates during OTC degradation was carried out. The feasibility of O3/CaO2/ HCO 3 for the treatment of other substances, such as bisphenol A, tetracycline, and actual wastewater, was investigated. Finally, the energy efficiency of the O3/CaO2/ HCO 3 system was calculated and compared with other mainstream processes of OTC degradation. The O3/CaO2/ HCO 3 system may be considered as an efficient and economical approach for antibiotic destruction.

1. Introduction

In recent decades, antibiotics have been widely used in the pharmaceutical and animal husbandry industries. While providing substantial social benefits, they inevitably cause serious environmental pollution [1,2]. Oxytetracycline (OTC) is a kind of tetracycline antibiotic [3] with broad-spectrum bacteriostasis and a low price. As a veterinary drug, it is widely used in animal husbandry worldwide [4]. However, excessive use of OTC cannot be fully absorbed by animals, and there will be a large amount of OTC in the environment [5,6]. A residual concentration of OTC in water is still present after treatment by sewage treatment plants. A high concentration of OTC in water will slow down the development of fish [7] and inhibit photosynthesis in plants [8]. The naphthalene ring structure has antibacterial properties and chemical stability, which makes it difficult to remove OTC through traditional water treatment processes [9]. Therefore, it is imperative to seek an effective method to remove OTC in water.
Oxidation is an essential reaction from an industrial and academic point of view. In general, selecting oxidants is significant for heterogeneous reaction systems. An advanced oxidation process (AOPS) is recognized as an effective method for treating wastewater with potential environmental risks. One of the most common active species in AOPS is hydroxyl radicals (·OH), which have strong oxidation properties and can convert pollutants into carbon dioxide and water. A simple and commonly used substance in AOPs is O3 [10]. O3 is considered to be a clean oxidant in which organic pollutants are oxidized directly or indirectly with free radicals generated by O3 to achieve the degradation of organic pollutants [11]. The direct oxidation reaction time is longer, the oxidation rate is slow, and it has a specific selectivity. Indirect oxidation refers to the solid oxidizing free radicals (mainly ·OH) generated by the catalytic O3 reaction, which can quickly decompose most organic pollutants in water. In the indirect oxidation method, the use of a catalyst to activate O3 increases the oxidation capacity of O3 and promotes the degradation of organic pollutants. Recently, the couple of O3 and other oxidation processes has become very popular, such as H2O2/O3 [12], ultrasonic treatment ozonation [13], and photocatalytic ozonation [14]. Among them, the most effective method of O3 indirect oxidation technology is the O3/H2O2 method, which is also the simplest method for ·OH [15]. However, H2O2 is relatively dangerous, with the characteristic of being flammable and explosive [16].
So far, many studies have confirmed that a combination of several treatment technologies can improve the degradation and mineralization of organic pollutants in wastewater [17]. Some catalysts can promote the better degradation of pollutants by O3. As a substitute for liquid hydrogen peroxide, calcium peroxide (CaO2), a new solid oxidant, has attracted a lot of attention in the environmental field due to its remarkable advantages, such as a high efficiency, safety, stability, controllability, and low cost [18,19]. The process of releasing H2O2 from calcium peroxide is slow, which can be controlled by adjusting the pH, which avoids the disadvantage of wasting oxygen and the short action time caused by the rapid decomposition of liquid H2O2. For some pollutants (such as PPCP and PAH), the removal effect of solid calcium peroxide will exceed that of liquid H2O2 [20,21]. The O3/H2O2 system created by O2/CaO2 has a low oxidation rate and poor selectivity in the absence of an activator. Sodium bicarbonate was added to the system as an auxiliary catalyst [22]. On one hand, HCO 3 reacts with H2O2 to generate peroxymonocarbonate ( HCO 4 ), which continuously activates H2O2 and converts it into highly active free radicals such as ·O2, singlet oxygen (1O2), and so on [23]. On the other hand, HCO 3 is a recognized promoter of O3 decomposition, promoting the efficient oxidation of O3 into ·CO3 and the conversion of O3 into indirect oxidation [24]. O3/CaO2/ HCO 3 innovatively proposes a safe and fast O3/H2O2 system. Meanwhile, the addition of HCO 3 could activate H2O2 and O3 and significantly improve the oxidation capacity of the system [25]. As a whole, the advantages of the O3/CaO2/ HCO 3 system can be listed as: firstly, the synergy effect between O3 and CaO2 can promote the generation of ·OH. Secondly, the OH released by CaO2 will form an alkaline environment, which will further promote the transformation of O3 into ·OH. Thirdly, HCO 3 activates O3 radicals and catalyzes the production of ·OH radicals. Finally, the decomposition products of CaO2/ HCO 3 are non-toxic and harmless. However, as far as is known, there have been few reports that O3 activates CaO2/ HCO 3 for antibiotic OTC degradation, and its synergistic mechanism is still unclear.
Therefore, the O3/CaO2/ HCO 3 system was constructed for the efficient degradation of OTC in this paper. The influence of the effect of CaO2 dosage, HCO 3 dosage, O3 concentration, initial pH value, and OTC initial concentration on OTC degradation was investigated. The role of active substances in the OTC decomposition system was inspected via the capture agent experiment. The generated active species were characterized using chemical methods and electron spin resonance (ESR). The degradation mechanism and process were examined. The toxicity of OTC and intermediates was analyzed, as was the feasibility of the O3/CaO2/ HCO 3 system for the treatment of other antibiotics and actual wastewater. Finally, the energy efficiency of the system was compared with that of other technology.

2. Results and Discussion

2.1. Effect of CaO2

Figure 1a,b reveal the changes in the OTC degradation efficiency under different CaO2 additions. The lowercase letters in the picture represent the difference between the data, and the star symbol represents the value of lnKobs. Experiments were carried out with fixed OTC concentrations at different CaO2 concentrations. It was observed that the O3/CaO2 system had a significant effect on the degradation of OTC within 30 min. When the dosage of CaO2 was 0.050 g/L, the degradation efficiency of OTC could reach 85.3%, which was 23.2% higher than that of the sole O3 system. The degradation efficiency (slope of degradation curve) of OTC increased with an increase in the CaO2 concentration, which indicates that CaO2 is conducive to the degradation of OTC. It is worth noting that, when the dosage of CaO2 was higher than the optimal dosage (0.050 g/L), the degradation efficiency gradually slowed down. The first-order kinetic constant in the upper right corner of Figure 1a can also reflect this trend.
CaO2 is a good solid source of H2O2 [26]. It can be regularly converted into H2O2 and O2 when dissolved in water (Equations (1) and (2)) [27]. The H2O2 it releases can reduce disproportionation and maintain the reaction for a longer time. H2O2 will cause O3 (E° = 2.07 V/NHE) to decompose and transform into non-selective ·OH (E° = 2.80 V/NHE) [28]. The formation of Ca(OH)2 will affect the pH value of the solution. The alkaline environment brought about makes H2O2 decompose into HO 2 . HO 2 is converted into ·OH through a series of reactions with O3 (Equations (3)–(13)) [29,30,31,32,33]. This ·OH is a strong oxidant, which can effectively promote the degradation of OTC. However, when the concentration of CaO2 is too high, too much H2O2 will be produced [34]. Excessive H2O2 will decompose by itself and need to consume active free radicals such as ·OH (Equations (14)–(16)), which will adversely affect the degradation of OTC [35]. The oxidation potentials of different free radicals are displayed in Table S1.
CaO 2 + 2 H 2 O     H 2 O 2 + Ca ( OH ) 2
H 2 O 2 + 2 O 3     2 OH + 3 O 2
OH + O 3     HO 2 + O 2
HO 2 + O 3     HO 2 + O 3
HO 2     · O 2 + · H
· O 3 + H +     · HO 3
· HO 3     O 2 + · OH
· O 3 + O 3   O 2 + · O 3
HO 2 + O 3   HO 5
HO 5     · HO 2 + · O 3
HO 5     2 O 2 + OH
· O 3     · O + O 2
· O + H 2 O     · OH + OH
2 H 2 O 2     O 2 + 2 H 2 O
H 2 O 2 + · OH     · HO 2 + H 2 O
HO 2 + · OH     · HO 2 + OH

2.2. Effect of H C O 3 - Dosage

The degradation of OTC with time under the conditions of different HCO 3 additions is shown in Figure 1c,d. To explore the CaO2/ HCO 3 combined process, with the addition of HCO 3 as a variable, the experiment was carried out with the optimal CaO2 concentration and fixed OTC concentration. It was observed that the O3/CaO2/ HCO 3 system had a further improvement in the degradation of OTC compared to O3/CaO2 within the reaction time of 30 min. When the dosage of HCO 3 was 2.25 mmol/L, the degradation efficiency of OTC could reach 91.5%, which was 7.3% higher than that of the O3/CaO2 system and 32.1% higher than that of the sole O3 system. The degradation efficiency of OTC increased with an increase in the HCO 3 concentration. Further increasing the concentration of HCO 3 based on the optimal dosage of HCO 3 will inhibit the degradation of OTC. Similarly, the first-order kinetic constant in the upper right corner of Figure 1c also reflects this trend.
HCO 3 reacts with H+ in water to produce CO2, H2O2 reacts with OH in water to produce OOH, and CO2 reacts with OOH to produce HCO 4 . The active free radicals released by HCO 4 ( · CO 3 ) can activate H2O2 to generate ·HO2, and then react to generate · O 2 , 1O2 and other oxygen-active substances (Equations (17)–(22)) [36,37]. The consumption of · CO 3 for activating H2O2 can be regenerated into · CO 3 through intermediate HCO 3 to form a cycle, which is the synergistic effect of HCO 3 and CaO2 ((Equation (19)) [23,38].
HCO 3 + H 2 O 2     HCO 4 + H 2 O
HCO 4     · CO 3 + · OH
HCO 3   +   · OH   CO 3   +   H 2 O
H 2 O 2 + · CO 3     HCO 3 + · HO 2
· HO 2     H +   +   · O 2
O 2 + · OH     O 1 2 + OH
The synergistic effect between HCO 3 and O3 is mainly reflected in: OH and · O 2 promote the decomposition of O3 into · O 3 , · O 3 , through a chain reaction, generates more · O 2 , · O 2 , in turn, promotes the decomposition of O3 (Equations (23)–(31)) [39,40,41]. There is a balance point between the generation of · O 2 and the decomposition of O3, which also explains the phenomenon that the degradation efficiency of OTC decreases when the HCO 3 dosing amount is higher than the optimal amount [42].
O 3 + · O 2     · O 3 + O 2
O 3   +   OH     H O 2 O 2
O 3   +   H O 2     · O 3   +   · HO 2
· O 3 + H +     · HO 3
· HO 3     · OH + O 2
O 3   +   · OH     · HO 2   +   O 2
H 2 O 2 + · OH     · HO 2   +   H 2 O
HO 2 + · OH     · O 2 + H 2 O  
· H O 2     · O 2   +   H +

2.3. Effect of O3 Dosage

The degradation of OTC under the conditions of different O3 concentrations is presented in Figure 1e,f. As the amount of O3 increased from 0.125 g/h to 0.75 g/h, the OTC degradation efficiency increased significantly from 91.5% to 97.2%, and the reaction time when the OTC degradation efficiency reached 90% was shortened from 30 min to 5 min. The reasons were as follows: firstly, the larger O3 flux increased the contact area between the gas phase and the liquid phase, and the number of effective molecular collisions increased, which improved the reaction rate. Secondly, according to Henry’s Law, the more significant the O3 concentration, that is, the increase in XZ per unit volume, the more increased the pressure of the O3 gas, thus promoting the mass transfer rate of O3 between the two phases [43]. Third, the O3 concentration will affect the concentrations of H2O2 and ·OH in the reaction system. A higher concentration will promote the generation of more active free radicals and affect the degradation effect [44]. However, a high-concentration O3 treatment process means a high cost, which will affect the economy of this process.

2.4. Effect pH Value

The degradation of OTC with time under the conditions of different initial pHs is shown in Figure 2a. With an increase in the initial pH of the solution, the environment changed from acidic to neutral and then to alkaline. The OTC degradation efficiency was enhanced from 71.4% to 94.4%. This phenomenon was mainly due to the direct oxidation of O3 molecules with pollutants in an acidic environment. In an alkaline environment, the concentration of dissolved O3 decreases [45], forcing O3 molecules to be activated into active free radicals such as ·OH and · O 2 (Equations (24)–(32)) to react with OTC [46]. Because the indirect reaction is dominant, the oxidation potential of ·OH is higher than that of O3 molecules, and the oxidation ability is stronger, which is more conducive to the degradation of OTC. On the other hand, an alkaline environment is more conducive to the release of H2O2 from CaO2 and the self-decomposition of H2O2, which will produce more active free radicals and promote the degradation of pollutants [47]. Similarly, the O3 treatment process under alkaline conditions also means an increase in cost, and the natural water environment is mainly neutral or weakly alkaline, so neutral environmental conditions will continue to be used in subsequent studies.

2.5. Effect OTC Concentration

The degradation of OTC with time under the conditions of different OTC concentrations is shown in Figure 2c. As the initial OTC concentration was raised from 40 mg/L to 200 mg/L, the OTC degradation efficiency decreased significantly, the degradation efficiency decreased from 94.4% to 60.6%, and the kinetic constant decreased from 0.114 min−1 to 0.031 min−1 (inset Figure 2c). It is unreasonable to judge the OTC degradation only by the initial OTC concentration. We calculated the total removal of the four initial OTC concentrations in Figure 2c, which were 15.1 mg, 29.0 mg, 45.8 mg, and 48.5 mg, respectively. Comparing the kinetic constant and total removal amount of OTC, with an increase in the concentration, the kinetic constant decreased and the total removal amount increased. Therefore, it is speculated that the initial OTC treatment concentration is 40 mg/L, which is close to the limit of the O3/CaO2/ HCO 3 system. As shown in Figure 1b,d,e and Figure 2b,d, good linear correlations between the ln k and different parameters were obtained. Therefore, the establishment of an O3/CaO2/ HCO 3 system can degrade OTC effectively.

2.6. Active Substance Analysis

2.6.1. Role of ·OH, 1O2, · O 2 , and · CO 3

Methanol is an efficient ·OH-trapping agent, which was used to capture ·OH in the O3/CaO2/ HCO 3 system. As shown in Figure 3a, it can be seen that the degradation efficiency of OTC decreased with increasing amounts of methanol. When the methanol concentration changed from 0 mmol/L to 10 mmol/L, the degradation efficiency of OTC decreased from 94.0% to 66.9%. Meanwhile, the kinetic constant declined from 0.114 min−1 to 0.039 min−1 (Figure S1a). These results indicate that ·OH played an essential role in the degradation of OTC. · O 2 can react to generate 1O2 (Equations (32)–(34)), which has a strong oxidation potential. 1,4-Diazabicyclooctane triethylenediamine (DABCO) is a highly efficient 1O2-trapping agent that can capture the 1O2 existing in the solution. Therefore, this section intends to use DABCO to study the role of 1O2 in the O3/CaO2/ HCO 3 system.
O 2 + · OH     O 1 2 + OH
· O 2 + HO 2 + H +     O 1 2   +   H 2 O 2
· HO 2 + HO 2     O 1 2 + H 2 O 2
Figure 3b shows the effect of various DABCO dosages on OTC elimination. Higher DABCO dosages will inhibit the degradation effect more obviously. When the addition of DABCO changed from 0 mmol/L to 10 mmol/L, the degradation efficiency of OTC decreased from 94.0% to 79.6%, with the kinetic constant varying from 0.114 min−1 to 0.060 min−1 (Figure S1b). DABCO has a significant inhibitory effect on OTC elimination. It is further inferred that 1O2 is actively involved in OTC degradation.
P-benzoquinone is a highly efficient · O 2 -trapping agent which can capture the · O 2 present in a solution. After adding the catalyst CaO2 to the solution, CaO2 and H2O will produce a large quantity of · O 2 (Equation (6)), which has a significant effect on OTC elimination. Figure 3c shows the effect of p-benzoquinone addition on OTC elimination. The degradation efficiency of OTC decreased with the p-benzoquinone addition. When the addition of p-benzoquinone was 0 mmol/L 0.5 mmol/L, 1.0 mmol/L, and 1.5 mmol/L, the degradation efficiency of OTC decreased from 94.0% to 53.2%, and the corresponding kinetic constant varied from 0.114 min−1 to 0.059 min−1, 0.039 min−1, and 0.033 min−1 (Figure S1c). This result showed that · O 2 played a key role in the degradation of OTC.
Indole is an efficient · CO 3 capture agent, which can capture the HCO 3 existing in a solution. After adding the catalyst NaHCO3 to the solution, NaHCO3 and H2O will produce part of · CO 3 (Equation (19)). When the indole concentration changed from 0 mmol/L to 0.2 mmol/L, the degradation efficiency of OTC decreased from 94.0% to 59.5% (Figure 3d), and the corresponding kinetic constant varied from 0.114 min−1 to 0.033 min−1 (Figure S1d). This result shows that · CO 3 is very involved in the degradation of OTC. The reaction rates of p-benzoquinone with ·OH and ·O2 were 1.2 × 109 and 8.3 × 108 M−1 s−1, respectively, which are comparable. The reaction rate of indole with ·OH was 7.9 × 10−14 M−1s−1, which can be ignored. Because of their low reaction rates, P-benzoquinone and Indole can be regarded as not reacting with ·OH [48,49]. As shown in Figure S2, compared with different active substances, ·O2 made the greatest contribution to the degradation process of OTC, with 41.2% of the OTC being degraded by ·O2.

2.6.2. Formation of Active Species

To further explore the synergetic mechanism, the formation of ·OH under different systems was compared. The results are depicted in Figure 4a. Compared to the sole O3 system, the ·OH concentration was seriously increased when CaO2 was added. These phenomena verified that O3 could react with the H2O2 released by CaO2 and promote the formation of ·OH [50]. As desired, the ·OH concentration was further enhanced when HCO 3 was added. The production of ·OH followed this law: O3/CaO2/NaHCO3 > O3/CaO2 > O3. Therefore, it can be explained that the higher degradation efficiency may have come from the higher production of ·OH. It is worth noting that the generation of ·OH in the O3/CaO2 system and the O3/CaO2/ HCO 3 system showed a trend of first rising, then decreasing, and lastly increasing over time. It was speculated that the conversion of H2O2 into ·OH was violent, and a large amount of ·OH was generated within 5 min. ·OH could react with H2O2, and a certain amount of ·OH production was consumed from 5 to 20 min. After the two reactions reached equilibrium, ·OH production increased and tended to stabilize.
ESR analyses further revealed that there were obvious signals of the ·OH peak in various systems (Figure 4b). It can be seen that the characterized peaks of ·OH followed as: O3/CaO2/ HCO 3 > O3/CaO2 > O3, which illustrated that ·OH was beneficial to generate in O3/CaO2/ HCO 3 system. Figure 4c presents the characteristic peaks of 1O2 in various systems. It can be seen that 1O2 can be generated in O3, O3/CaO2, and O3/CaO2/ HCO 3 systems. Compared to sole O3, the characteristic peaks of 1O2 were enhanced with the addition of CaO2 and HCO 3 , which indicates that the addition of CaO2 and HCO 3 effectively increased the production of 1O2.

2.7. OTC Degradation Process

2.7.1. UV-Vis Spectra

Figure S3 shows the UV-Vis spectrum results of the OTC solution in the sole O3 system and O3/CaO2/ HCO 3 system. The OTC spectrum shows two absorption bands near 276 nm and 353 nm, and the absorption peak near 353 nm is the characteristic peak of OTC. The common point of the two systems is that, after 30 min of processing, the characteristic peaks of OTC were significantly reduced, and the OTC molecules were weakened. The π–π * transition of its heteroaromatic ring makes the main contribution to this [51,52]. Compared with Figure S3a, the characteristic peak of OTC in Figure S3b has a more rapid decline, which indicates that O3/CaO2/ HCO 3 had a better OTC degradation ability.

2.7.2. 3D EEMFS

To study the degradation process of OTC more deeply, 3D EEMFs were used for characterization. The emission wavelength and the excitation wavelength were all set as 200–600 nm, whose sampling intervals were 5 nm and 10 nm, respectively. Figure 5a is taken from the original OTC sample, and Figure 5b,c are OTC samples processed for 10 min, 20 min, and 30 min, respectively. In the original OTC solution, there was a prominent fluorescence peak at Ex/Em = 300–400/400–600 nm, which is located in the humic acid-like fluorescence region [53]. With the passage of time, the characteristic peak intensity gradually weakened and the fluorescence intensity dropped from 2000.23 to 220.18, indicating that the OTC molecule was destroyed and may have been mineralized into some intermediates or decomposed into CO2 and H2O (Figure 5d). The 3D EEMFs results show that the O3/CaO2/ HCO 3 system can degrade OTC, which is consistent with the UV-Vis spectrum results.

2.7.3. Variation of TOC and COD

TOC is an important parameter for evaluating the mineralization ability of oxidation systems. Figure 6a shows the degradation of OTC in terms of TOC removal, with or without two catalysts in the O3 oxidation system. The carbon-containing organic matter in the OTC solution decreased as the treatment time increased. After the two catalysts were added, the TOC concentration decreased. When the treatment time was 30 min, the TOC degradation efficiency could reach 14.7%. Figure 6b depicts the COD concentration in the two systems. The COD concentration of the sole O3 system underwent a slight drop, the speed was plodding, and it tended to be stable. The O3/CaO2/ HCO 3 system had a very high COD concentration at 0 min, with a significant drop and a fast speed, and a COD degradation efficiency of 48.8% was obtained within 30 min of the processing time. This O3/CaO2/ HCO 3 system produced a large number of active free radicals, which actively participated in the degradation process, which supports the derivation of the mechanism in the previous chapter. The O3/CaO2/ HCO 3 system has an excellent OTC degradation ability, but the degradation efficiency of TOC is far below 100%. Combining the results of the UV-Vis spectroscopy and 3D EEMFs, we speculated that a great deal of intermediates were formed during OTC degradation.

2.7.4. Variation of pH and Conductivity

Figure S4a shows the pH change over time during OTC degradation. The O3/CaO2/ HCO 3 system presented a slightly alkaline environment, while the sole O3 system presented a neutral environment. As the treatment time increased, the pH values in both systems showed a gentle downward trend. Figure S4b shows the change in conductivity during OTC degradation. The conductivity of the sole O3 system was 21.2 μS/cm at 0 min, and after 30 min of treatment, the conductivity increased slightly to 30.3 μS/cm. The conductivity of the O3/CaO2/ HCO 3 system was 308 μS/cm at 0 min and dropped to 262 μS/cm after 30 min of treatment. We speculate that the addition of catalysts produced more inorganic ions, and acidic free radicals were generated during the degradation of OTC. In the process of degradation, HCO 3 reacted with H2O2 to form HCO 4 on the one hand and ·OH to form · CO 3 on the other, so HCO 3 itself was consumed massively. The decrease in conductivity may have been due to the consumption of HCO 3 .

2.8. DFT Analysis and Degradation Pathway

The Fukui functions calculated with density functional and electron orbitals provide good evidence for predicting reaction sites. The molecular structure of OTC was calculated with the Gaussian 09 program. The Fukui function of OTC mapped the electron density isosurface and the calculation results are provided in Figure 7a and Table S2. Now, it is generally accepted that, the higher the f+ function of the atom is, the easier will it be nucleophilically attacked; the higher the f function of the atom is, the easier will it be electrophilically attacked; and the higher the f0 function of the atom is, the easier will it be radically attacked. To indicate the possibility of the site being attacked clearly based on the above mechanism, electron cloud diagrams were drawn. It can be seen that the highlight is above C21 and O26, which means that these two sites would be easy to electrophilically attack. For the same reason, C4, O12, and C20 were easier to nucleophilically attack. C21 and O26 were easier to radically attack. Through a comprehensive comparison of the Fukui function table, it is inferred that OTC molecules may be decomposed into A (pathway 1) and B (pathway 2) due to electron and free radical attacks on C21 and O26.
To deeply explore the OTC degradation process in the O3/CaO2/ HCO 3 system, a Liquid Chromatograph Mass Spectrometer (LC-MS) (ES+, ES−) was adopted and the results are illustrated in Figure S5. According to the results of LC-MS, it can be inferred that there were eight main intermediates: A C22H24N2O10 (m/z = 477), B C15H13NO7 (m/z = 296), C C4H7NO4 (m/z = 132), D C3H7NO4 (m/z = 118), E C3H7NO3 (m/z = 106), F C6H10O2 (m/z = 113), and G C6H12O (m/z = 101). The molecular structures of the eight intermediates are shown in Table S3. Combined with the Density Functional Theory (DFT) calculation and LC-MS detection results, the inferred degradation path of OTC is shown in Figure 7b. Figure 7c shows a free energy ladder diagram of these reactions. The initial energy of OTC was 0 kcal/mol, and the major degradation pathways are represented by negative energy levels. These reactions may occur, in theory, by different radicals. In the OTC molecular structure formula, weak chemical bonds between atoms are easy to destroy. C21 in pathway I is prone to enol ketone isomerization, and OTC is transformed into more stable isomers. The ketone/enol portion of C21–C25 is vulnerable to ·OH attack, and further hydroxylation by-product A is obtained. O3 in the system is an electrophilic reagent, which tends to attack the electron-rich region in the OTC structure [54]. The double-bond structure of C21–C25 in pathway II and pathway III is broken by free radicals such as O3, and the main product is B. Fragile C5a breaks, causing B molecules to decompose into C and F. F is further dehydrated to obtain by-product G. C is demethylated to obtain by-product D, and D is further dehydrated to obtain small molecule by-product E. These intermediates could be further oxidized to small molecular organic acids, and ultimately mineralized to CO2 and H2O [55,56].

2.9. Toxicity Evaluation

The U.S. Environmental Protection Agency’s toxicity evaluation software tool (US–EPA–TEST, version 5.1.2) program of this study predicts the acute and chronic toxicity of OTC and its intermediates [55,57]. The European Union criteria standard is adopted for acute toxicity, and the Chinese guidelines for the evaluation of hazardous chemicals (HJ/TI 154—2004) are adopted for chronic toxicity. The test results are shown in Table S4 and Figure 8. The daphnid LC50 of OTC is 9.34 mg/L, belonging to “toxic”. After 30 min of treatment in this system, the toxicity of intermediates was significantly reduced, and all intermediates except F were located in “harmless” or “harmful” areas. A, B, and C were located in the “Harmless” area. Therefore, it can be speculated that the toxicity of OTC will decline significantly after O3/CaO2/ HCO 3 treatment, and detoxification may be achieved.

2.10. Feasible for Treatment of Other Antibiotics and Actual Wastewater

The feasibility of O3/CaO2/ HCO 3 for the treatment of other antibiotics and actual wastewater was investigated. The degradation of bisphenol A (BPA) and tetracycline (TC) under different systems is presented in Figure 9a,b. No matter the aspect of BPA or TC, the highest degradation efficiency was obtained in the O3/CaO2/ HCO 3 system. As a whole, it can be seen that the degradation efficiency could reach above 80%, illustrating that the O3/CaO2/ HCO 3 system is also adapted to other antibiotic degradations. Figure 9c depicts the degradation of OTC in various water bodies. Over time, OTC molecules are gradually degraded in different water bodies such as deionized water, lake water, river water, and tap water. The degradation effect has not undergone significant changes, indicating that the system is also suitable for the treatment of actual wastewater.

2.11. Energy Efficiency Evaluation

Energy efficiency assessment is a part that cannot be underestimated. It is closely related to whether a new system can be promoted and applied. Therefore, we calculated the energy efficiency of the sole O3 system, O3/CaO2 system, and O3/CaO2/ HCO 3 system in different program doses, and the results are shown in Table S5. In the sole O3 system, the energy efficiency reached 1.55 g/kWh. When 0.050 g/L of CaO2 was added, the energy efficiency increased to 1.82 g/kWh. On this basis, when adding 2.25 mmol/L of HCO 3 , the optimal energy efficiency could be increased to 1.91 g/kWh. The energy efficiency of the optimal O3/CaO2/ HCO 3 system could reach 1.91 g/kWh, which is an impressive result.
Table S6 lists the comparison results of the kinetic constants and energy efficiencies of other degradation OTC processes and this process. The degradation efficiency of the Al0-Gr-Fe0/O2 Fenton process for OTC was higher than that of other systems, close to 100%, but the energy efficiency was extremely low [40]. CO3O4/CNT catalyzed the degradation of OTC, and its degradation efficiency was close to O3/CaO2/ HCO 3 , but required a longer processing time and a relatively small kinetic constant [43]. Among the many degradation OTC processes, O3/CaO2/ HCO 3 is undoubtedly the most economical and has the most potential [41,42,44,45].

3. Materials and Methods

3.1. Chemicals

OTC, CaO2, NaOH, triethylenediamine, terephthalic acid, sodium indigo disulfonate, BPA, and TMPO were all analytical grade (AR), and were bought from Aladdin Reagent (Aladdin Reagent (Shanghai) Co., Ltd., Shanghai, China). H2SO4 (98%) and H3PO4 were purchased from Sinopharm Chemical Reagent (Sinopharmaceutical Group Chemical Reagent (Shanghai) Co., Ltd., Shanghai, China). Indole, TC, and DMPO were obtained from Macklin (McLean biochemical Technology (Shanghai) Co., Ltd., Shanghai, China). NaHCO3 of analytical grade (AR). was obtained from Lingfeng Chemical Reagent (Lingfeng Chemical Reagent (Shanghai) Co., Ltd., Shanghai, China). None of the chemicals required further purification. All of the solution was made with deionized water produced by an ultra-pure water system (Biosafer, Biosafer—10R, Saifei (Nanjing) Co., Ltd., Nanjing, China).

3.2. Experimental Process

A reactor schematic diagram is presented in Figure S6. The O3 was generated by plasma discharge, which was named as plasma-O3 generator (CQ, Changqing—802S). The amount of O3 used was extracted and adjusted using an atmospheric sampler. The active species were characterized using an electron paramagnetic resonance spectrometer (BRUKER, MESRV30077, Brooke Technology (Beijing) Co., Ltd., Beijing, China). To monitor the OTC degradation process, 3 mL of the aqueous solution was taken out of the reaction mixture every 5 min, and the experiment was repeated several times to take the average value for the subsequent analysis. The best degradation conditions were as follows: the concentration of CaO2 was 0.050 g/L, the dosage of O3 was 0.75 g/h, the dosage of HCO 3 was 2.25 mmol/L, the pH was 2.4, and the initial concentration of OTC was 40 mg/L.

3.3. Analysis

The OTC calculation and analysis process is provided in Text S1. For the detection of O3, ·OH, and H2O2 in the OTC degradation process, the indigo method [58], the terephthalic acid probe method [59], and the titanyl sulfate method [60] were used, respectively. The data were analyzed and plotted with the origin (OriginLab Inc., version 9.8.0.200, Northampton, MA, USA) software.

3.4. DFT Analysis and Toxicity Evaluation

The molecular structures of the organic compounds studied were built using the Gauss View 6.0 software (Gaussian Inc., Wallingford, CT, USA). The built molecular structures were then optimized using chemical DFT methods (B3LYP/6-311++G (d, p)), employing Gaussian 16 (Gaussian Inc., Wallingford, CT, USA). Detailed information is shown in Text S2. US-EPA-TEST (US EPA, version 5.1.2, USA) was adopted to assess the toxicity of OTC and its intermediates based on the bioaccumulation factor and LC50 [57].

4. Conclusions

This study constructed O3/CaO2/ HCO 3 oxidation technology to degrade OTC wastewater. The introduction of CaO2/ HCO 3 significantly promoted the degradation of OTC in the O3 process in terms of degradation efficiency and kinetic constants. However, an excessive CaO2/ HCO 3 dosage had an inhibitory effect on OTC degradation. The greater the O3 flow rate, the lower the OTC concentration, and the higher the pH, which were beneficial for OTC decomposition. Active substances such as ·OH, · O 2 , 1O2, and · CO 3 participated in the degradation of OTC. The enhanced degradation of OTC was mainly due to the high production of ·OH. After the treatment of the O3/CaO2/ HCO 3 system, TOC and COD had both declined. However, as the reaction time increased, the pH value was reduced, with or without catalysts. The conductivity was enhanced without a catalyst, while an opposite trend was observed for the O3/CaO2/ HCO 3 system. According to a DFT analysis, LC-MS, and related studies, the degradation pathways related to OTC degradation were proposed. After the O3/CaO2/ HCO 3 system treatment, the toxicity of OTC could be reduced. O3/CaO2/ HCO 3 is also feasible for the treatment of other antibiotics and actual wastewater. The energy efficiencies for OTC degradation under various systems were compared. This research showed that the O3/CaO2/ HCO 3 system could be considered as an efficient and commercial oxidation process for OTC treatment.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29030659/s1. Figure S1. First-order kinetic fitting and kinetic constants of scavenger: (a) methanol; (b) DABCO; (c) p-benzoquinone; and (d) indole. Figure S2. Contribution rates of different free radicals. Figure S3. UV-Vis curve of OTC degradation in (a) O3 system and (b) O3/CaO2/ HCO 3 system. Figure S4. Variation in (a) pH and (b) conductivity during OTC degradation. Figure S5. MS spectra of OTC degradation intermediates identified by LC-MS. Figure S6. Schematic representation of the O3 reactor for the oxidative degradation of OTC. Table S1. Oxidation potential of active radical species in O3/CaO2/ HCO 3 system. Table S2. Fukui function of OTC. Table S3. The proposed structure information of OTC degradation products. Table S4. Toxicity of OTC degradation intermediates. Table S5. Energy efficiency during OTC elimination with various CaO2 and NaHCO3 dosages. Table S6. Comparison with other techniques on OTC elimination. References [40,41,42,43,44,45] are cited in the supplementary materials.

Author Contributions

Conceptualization, D.Z. and H.G.; methodology, L.X. and S.P.; formal analysis, Z.L. and H.G.; funding acquisition, S.L.; investigation, L.X. and S.P.; data curation, Z.L.; writing—original draft preparation, Z.L and L.X.; writing—review and editing, H.G.; visualization, S.L.; supervision, H.G.; project administration, H.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (No. 22006069), and Natural Science Foundation of Jiangsu Province in China (No. BK20200801).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

Author Shen Li is employed by the company Anhui Jiuwu Tianhong Environmental Protection Technology Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Boyd, N.; Teng, C.; Frei, C. Brief Overview of approaches and challenges in new antibiotic development: A focus on drug repurposing. Front. Cell. Infect. Microbiol. 2021, 17, 684515. [Google Scholar] [CrossRef] [PubMed]
  2. Sodhi, K.; Kumar, M.; Balan, B.; Dhaulaniya, A.; Shree, P.; Sharma, N.; Singh, D. Perspectives on the antibiotic contamination, resistance, metabolomics, and systemic remediation. SN Appl. Sci. 2021, 3, 269. [Google Scholar] [CrossRef]
  3. Guo, H.; Wang, Y.; Yao, X.; Zhang, Y.; Li, Z.; Pan, S.; Han, J.; Xu, L.; Qiao, W.; Li, J.; et al. A comprehensive insight into plasma-catalytic removal of antibiotic oxytetracycline based on graphene-TiO2-Fe3O4 nanocomposites. Chem. Eng. J. 2021, 425, 130614. [Google Scholar] [CrossRef]
  4. Feng, Y.; Wang, G.; Liu, Y.; Cheng, D.; Fan, S.; Zhao, Q.; Xue, J.; Zhang, S.; Li, Z. The impacts of oxytetracycline on humification during manure composting can be alleviated by adjusting initial moisture contents as illustrated by NMR. J. Integr. Agric. 2021, 20, 2277–2288. [Google Scholar] [CrossRef]
  5. Yang, Y.; Song, W.; Lin, H.; Wang, W.; Du, L.; Xing, W. Antibiotics and antibiotic resistance genes in global lakes: A review and meta-analysis. Environ. Int. 2018, 116, 60–73. [Google Scholar] [CrossRef] [PubMed]
  6. Qi, W.; Long, J.; Feng, C.; Feng, Y.; Cheng, D.; Liu, Y.; Xue, J.; Li, Z. Fe3+ enhanced degradation of oxytetracycline in water by pseudomonas. Water Res. 2019, 160, 361–370. [Google Scholar] [CrossRef]
  7. Liu, S.; Zhao, H.; Lehmler, H.; Cai, X.; Chen, J. Antibiotic pollution in marine food webs in laizhou bay, north china: Trophodynamics and human exposure implication. Environ. Sci. Technol. 2017, 51, 2392–2400. [Google Scholar] [CrossRef]
  8. Miguel, G.; Soledad, G.; Ismael, R.; Francisco, L.; Roberto, R.; Karina, B.; Eduardo, M.; Francisca, F. Toxicity of five antibiotics and their mixtures towards photosynthetic aquatic organisms: Implications for environmental risk assessment. Water Res. 2013, 47, 2050–2064. [Google Scholar]
  9. Liu, Y.; He, X.; Duan, X.; Fu, Y.; Dionysiou, D. Photochemical degradation of oxytetracycline: Influence of pH and role of carbonate radical. Chem. Eng. J. 2015, 276, 113–121. [Google Scholar] [CrossRef]
  10. Eghbali, P.; Hassani, A.; Wacławek, S.; Lin, K.A.; Sayyar, Z.; Ghanari, F. Recent advances in design and engineering of MXene-based catalysts for photocatalysis and persulfate-based advanced oxidation processes: A state-of-the-art review. Chem. Eng. J. 2024, 480, 147920. [Google Scholar] [CrossRef]
  11. Malik, S.; Ghosh, P.; Vaidya, A.; Mudliar, S. Hybrid ozonation process for industrial wastewater treatment: Principles and applications: A review. J. Water Process. Eng. 2020, 35, 101193. [Google Scholar] [CrossRef]
  12. Mohammad, A.; Aurelio, H. Reducing the formation of trihalomethanes (THMs) by ozone combined with hydrogen peroxide (H2O2/O3). Desalination 2006, 194, 121–126. [Google Scholar]
  13. Wu, Z.; Abramova, A.; Nikonov, R.; Cravotto, G. Sonozonation (sonication/ozonation) for the degradation of organic contaminants-A review. Ultrason. Sonochem. 2020, 68, 105195. [Google Scholar] [CrossRef] [PubMed]
  14. Mecha, A.C.; Chollom, M.N. Photocatalytic ozonation of wastewater: A review. Environ. Chem. Lett. 2020, 18, 1491–1507. [Google Scholar] [CrossRef]
  15. Anu, M.; Mika, S. Removal of natural organic matter from drinking water by advanced oxidation processes. Chemosphere 2010, 80, 351–365. [Google Scholar]
  16. Xue, Y.; Rajic, L.; Chen, L.; Lyu, S.; Alshawabkeh, A. Electrolytic control of hydrogen peroxide release from calcium peroxide in aqueous solution. Electrochem. Commun. 2018, 93, 81–85. [Google Scholar] [CrossRef]
  17. Pelalak, R.; Hassani, A.; Heidari, Z.; Zhou, M. State-of-the-art recent applications of layered double hydroxides (LDHs) material in Fenton-based oxidation processes for water and wastewater treatment. Chem. Eng. J. 2023, 474, 145511. [Google Scholar] [CrossRef]
  18. Xu, Q.; Huang, Q.S.; Wei, W.; Sun, J.; Dai, X.; Ni, B.J. Improving the treatment of waste activated sludge using calcium peroxide. Water Res. 2020, 187, 116440. [Google Scholar] [CrossRef]
  19. Yuan, D.; Zhang, C.; Tang, S.; Wang, Z.; Sun, Q.; Zhang, X.; Jiao, T.; Zhang, Q. Ferric ion-ascorbic acid complex catalyzed calcium peroxide for organic wastewater treatment: Optimized by response surface method. Chin. Chem. Lett. 2021, 32, 3387–3392. [Google Scholar] [CrossRef]
  20. Zhang, A.; Wang, J.; Li, Y. Performance of calcium peroxide for removal of endocrine-disrupting compounds in waste activated sludge and promotion of sludge solubilization. Water Res. 2015, 71, 125–139. [Google Scholar] [CrossRef]
  21. Zhang, A.; Zhou, Y.; Li, Y.; Liu, Y.; Li, X.; Xue, G.; Miruka, A.C.; Zheng, M.; Liu, Y. Motivation of reactive oxygen and nitrogen species by a novel non-thermal plasma coupled with calcium peroxide system for synergistic removal of sulfamethoxazole in waste activated sludge. Water Res. 2022, 212, 118128. [Google Scholar] [CrossRef] [PubMed]
  22. Fang, Y.; Yang, Z.; Zhang, X.; Ji, H. Synergistic catalytic oxidation of cinnamaldehydes by poly (vinyl alcohol) functionalized β-cyclodextrin polymer in CaO2/ HCO 3 system. Supramol. Chem. 2018, 30, 134–145. [Google Scholar] [CrossRef]
  23. Li, Y.; Li, L.; Chen, Z.; Zhang, J.; Gong, L.; Wang, Y.; Zhao, H.; Mu, Y. Carbonate-activated hydrogen peroxide oxidation process for azo dye decolorization: Process, kinetics, and mechanisms. Chemosphere 2018, 192, 372–378. [Google Scholar] [CrossRef] [PubMed]
  24. Garoma, T.; Umamaheshwar, S.; Mumper, A. Removal of sulfadiazine, sulfamethizole, sulfamethoxazole, and sulfathiazole from aqueous solution by ozonation. Chemosphere 2010, 79, 814–820. [Google Scholar] [CrossRef] [PubMed]
  25. Yang, R.; Zeng, G.; Xu, Z.; Zhou, Z.; Huang, J.; Fu, R.; Lyu, S. Comparison of naphthalene removal performance using H2O2, sodium percarbonate and calcium peroxide oxidants activated by ferrous ions and degradation mechanism. Chemosphere 2021, 283, 131209. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, H.; Shen, Z.; Yan, X.; Guo, H.; Mao, D.; Yi, C. Dielectric barrier discharge plasma coupled with WO3 for bisphenol A degradation. Chemosphere 2021, 274, 129722. [Google Scholar] [CrossRef]
  27. Wang, X.; Wang, P.; Liu, X.; Hu, L.; Wang, Q.; Xu, P.; Zhang, G. Enhanced degradation of PFOA in water by dielectric barrier discharge plasma in a coaxial cylindrical structure with the assistance of peroxymonosulfate. Chem. Eng. J. 2020, 389, 124381. [Google Scholar] [CrossRef]
  28. Tang, C.; Zhang, Y.; Han, J.; Tian, Z.; Chen, L.; Chen, J. Monitoring graphene oxide’s efficiency for removing Re(VII) and Cr(VI) with fluorescent silica hydrogels. Environ. Pollut. 2020, 262, 11426. [Google Scholar] [CrossRef]
  29. Song, S.; Xu, X.; Xu, L.; He, Z.; Ying, H.; Chen, J. Mineralization of CI Reactive Yellow 145 in Aqueous Solution by UItraviolet-Enhanced Ozonation. Ind. Eng. Chem. Res. 2008, 47, 1386–1391. [Google Scholar] [CrossRef]
  30. Qian, Y.; Zhang, J.; Zhang, Y.; Chen, J.; Zhou, X. Degradation of 2,4-dichlorophenol by nanoscale calcium peroxide: Implication for groundwater remediation. Sep. Purif. Technol. 2016, 166, 222–229. [Google Scholar] [CrossRef]
  31. Gunten, U. Ozonation of drinking water: Part II Disinfection and by-product formation in presence of bromide, iodide or chlorine. Water Res. 2003, 37, 1479–1487. [Google Scholar] [CrossRef]
  32. Muhammad, A.; Abdul, G.L.; Salman, A. Water purification by electrical discharges, Plasma sources. Sci. Technol. 2001, 10, 82–91. [Google Scholar]
  33. Merényi, G.; Lind, J.; Naumov, S.; Sonntag, C. Reaction of ozone with hydrogen peroxide (peroxone process): A revision of current mechanistic concepts based on thermokinetic and quantum-chemical considerations. Environ. Sci. Technol. 2010, 44, 3505–3507. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, Y.; Qu, G.; Sun, Q.; Jia, H.; Wang, T.; Zhu, L. Endogenously activated persulfate by non-thermal plasma for Cu(II)-EDTA decomplexation: Synergistic effect and mechanisms. Chem. Eng. J. 2021, 406, 126774. [Google Scholar] [CrossRef]
  35. Chen, Y.; Duan, X.; Zhou, X.; Wang, R.; Wang, S.; Ren, N.; Ho, S. Advanced oxidation processes for water disinfection: Features, mechanisms and prospects. Chem. Eng. J. 2021, 409, 128207. [Google Scholar] [CrossRef]
  36. Shang, K.; Li, W.; Wang, X.; Lu, N.; Jiang, N.; Li, J.; Wu, Y. Degradation of p-nitrophenol by DBD plasma/Fe2+/persulfate oxidation process. Sep. Purifi. Technol. 2019, 218, 106–112. [Google Scholar] [CrossRef]
  37. Fu, X.; Gu, X.; Lu, S.; Miao, Z.; Xu, M.; Zhang, X.; Qiu, Z.; Sui, Q. Benzene depletion by Fe2+-catalyzed sodium percarbonate in aqueous solution. Chem. Eng. J. 2015, 267, 25–33. [Google Scholar] [CrossRef]
  38. Fu, X.; Gu, X.; Lu, S.; Xu, M.; Miao, Z.; Zhang, X.; Zhang, Y.; Xue, Y.; Qiu, Z.; Sui, Q. Enhanced degradation of benzene in aqueous solution by sodium percarbonate activated with chelated-Fe(II). Chem. Eng. J. 2016, 285, 180–188. [Google Scholar] [CrossRef]
  39. Jawad, A.; Lu, X.; Chen, Z.; Yin, G. Degradation of chlorophenols by supported Co-Mg-Al layered double hydrotalcite with bicarbonate activated hydrogen peroxide. J. Phys. Chem. A 2014, 118, 10028–10035. [Google Scholar] [CrossRef]
  40. Liu, Y.; Fan, Q.; Wang, J. Zn-Fe-CNTs catalytic in situ generation of H2O2 for Fenton-like degradation of sulfamethoxazole. J. Hazard. Mater. 2018, 342, 166–176. [Google Scholar] [CrossRef]
  41. Ni, J.; Liu, D.; Wang, W.; Wang, A.; Jia, J.; Tian, J.; Xing, Z. Hierarchical defect-rich flower-like BiOBr/Ag nanoparticles/ultrathin g-C3N4 with transfer channels plasmonic Z-scheme heterojunction photocatalyst for accelerated visible-light-driven photothermal-photocatalytic oxytetracycline degradation. Chem. Eng. J. 2021, 419, 129969. [Google Scholar] [CrossRef]
  42. Ma, W.; Yao, B.; Zhang, W.; He, Y.; Yu, Y.; Niu, J. Fabrication of PVDF-based piezocatalytic active membrane with enhanced oxytetracycline degradation efficiency through embedding few-layer E-MoS2 nanosheets. Chem. Eng. J. 2021, 415, 129000. [Google Scholar] [CrossRef]
  43. Liu, D.; Li, M.; Li, X.; Ren, F.; Sun, P.; Zhou, L. Core-shell Zn/Co MOFs derived Co3O4/CNTs as an efficient magnetic heteroge-neous catalyst for persulfate activation and oxytetracycline degradation. Chem. Eng. J. 2020, 387, 124008. [Google Scholar] [CrossRef]
  44. Brown, K.; Kulis, J.; Thomson, B.; Chapman, T.; Mawhinney, D. Occurrence of antibiotics in hospital, residential, and dairy effluent, municipal wastewater, and the rio grande in new Mexico. Sci. Total Environ. 2006, 366, 772–783. [Google Scholar] [CrossRef] [PubMed]
  45. Wang, L.; Yang, H.; Zhang, C.; Mo, Y.; Lu, H. 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]
  46. Back, J.; Obholzer, T.; Winkler, K.; Jabornig, S.; Rupprich, M. Combining ultrafiltration and non-thermal plasma for low energy degradation of pharmaceuticals from conventionally treated wastewater. J. Environ. Chem. Eng. 2018, 6, 7377–7385. [Google Scholar] [CrossRef]
  47. Han, Q.; Dong, W.; Wang, H.; Ma, H.; Liu, P.; Gu, Y.; Fan, H.; Song, X. Degradation of tetrabromobisphenol a by ozonation: Performance, products, mechanism and toxicity. Chemosphere 2019, 235, 701–712. [Google Scholar] [CrossRef] [PubMed]
  48. Guo, Y.; Zhang, Y.; Yu, G.; Wang, Y. Revisiting the role of reactive oxygen species for pollutant abatement during T catalytic ozonation: The probe approach versus the scavenger approach. Appl. Catal. B Environ. 2021, 280, 119418. [Google Scholar] [CrossRef]
  49. Xue, J.; Ma, F.; Elm, J.; Chen, J.; Xie, H. Atmospheric oxidation mechanism and kinetics of indole initiated by ·OH and ·Cl: A computational study. Atmos. Chem. Phys. 2022, 22, 11543–11555. [Google Scholar] [CrossRef]
  50. Sun, J.; Shen, C.; Guo, J.; Guo, H.; Yin, Y.; Xu, X.; Fei, Z.; Liu, Z.; Wen, X. Highly efficient activation of peroxymonosulfate by CO3O4/Bi2WO6 p-nheterojunction composites for the degradation of ciprofloxacin under visible light irradiation. J. Colloid Interface Sci. 2021, 588, 19–30. [Google Scholar] [CrossRef]
  51. Chen, H.; Wang, J. Degradation and mineralization of ofloxacin by ozonation and peroxone (O3/H2O2) process. Chemosphere 2021, 269, 128775. [Google Scholar] [CrossRef] [PubMed]
  52. Sun, K.; Yuan, D.; Liu, Y.; Song, Y.; Sun, Z.; Liu, R. Study on the efficiency and mechanism of direct red 80 dye by conventional ozonation and peroxone (O3/H2O2) treatment. Sep. Sci. Technol. 2020, 55, 3175–3183. [Google Scholar] [CrossRef]
  53. Radko, M.; Rutkowska, M.; Kowalczyk, A.; Mikrut, P.; Swięs, A.; Díaz, U.; Palomares, A.; Macyk, W.; Chmielarz, L. Catalytic oxidation of organic sulfides by H2O2 in the presence of titanosilicate zeolites. Microporous Mesoporous Mater. 2020, 302, 110219. [Google Scholar] [CrossRef]
  54. Khan, K.; Bae, H.; Jung, J. Tetracycline degradation by ozonation in the aqueous phase: Proposed degradation intermediates and pathway. J. Hazard. Mater. 2010, 181, 659–665. [Google Scholar] [CrossRef] [PubMed]
  55. Zhang, T.; Zhou, R.; Wang, P.; Mai-Prochnow, A.; Mcconchie, R.; Li, W.; Zhou, R.; Thompson, E.W.; Ostrikov, K.; Cullen, P.J. Degradation of cefixime antibiotic in water by atmospheric plasma bubbles: Performance, degradation pathways and toxicity evaluation. Chem. Eng. J. 2021, 421, 127730. [Google Scholar] [CrossRef]
  56. Han, S.; Mao, D.; Wang, H.; Guo, H. An insightful analysis of dimethyl phthalate degradation by the collaborative process of DBD plasma and Graphene-WO3 nanocomposites. Chemosphere 2022, 291, 132774. [Google Scholar] [CrossRef] [PubMed]
  57. Guo, H.; Li, Z.; Lin, S.; Li, D.; Jiang, N.; Wang, H.; Han, J.; Li, J. Multi-catalysis induced by pulsed discharge plasma coupled with graphene-Fe3O4 nanocomposites for efficient removal of ofloxacin in water: Mechanism, degradation pathway and potential toxicity. Chemosphere 2021, 265, 129089. [Google Scholar] [CrossRef]
  58. Guo, H.; Jiang, N.; Wang, H.; Shang, K.; Lu, N.; Li, J.; Wu, Y. Enhanced catalytic performance of graphene-TiO2 nanocomposites for synergetic degradation of fluoroquinolone antibiotic in pulsed discharge plasma system. Appl. Catal. B Environ. 2019, 248, 552–566. [Google Scholar] [CrossRef]
  59. Wang, T.; Jia, H.; Guo, X.; Xia, T.; Qu, G.; Sun, Q.; Yin, X. Evaluation of the potential of dimethyl phthalate degradation in aqueous using sodium percarbonate activated by discharge plasma. Chem. Eng. J. 2018, 346, 65–76. [Google Scholar] [CrossRef]
  60. Guo, H.; Li, Z.; Zhang, Y.; Jiang, N.; Wang, H.; Li, J. Degradation of chloramphenicol by pulsed discharge plasma with heterogeneous Fenton process using Fe3O4 nanocomposites. Sep. Purif. Technol. 2020, 253, 117540. [Google Scholar] [CrossRef]
Figure 1. Impact of CaO2 addition: (a) degradation efficiency and (b) linear fitting of dynamic constants; impact of NaHCO3 addition: (c) degradation efficiency and (d) linear fitting of dynamic constants; impact of O3 dosage: (e) degradation efficiency and (f) linear fitting of dynamic constants. Experimental conditions: [OTC] = 80 mg/L, 400 mL, 32 mg.
Figure 1. Impact of CaO2 addition: (a) degradation efficiency and (b) linear fitting of dynamic constants; impact of NaHCO3 addition: (c) degradation efficiency and (d) linear fitting of dynamic constants; impact of O3 dosage: (e) degradation efficiency and (f) linear fitting of dynamic constants. Experimental conditions: [OTC] = 80 mg/L, 400 mL, 32 mg.
Molecules 29 00659 g001
Figure 2. Impact of pH value: (a) degradation efficiency; (b) linear fitting of dynamic constants, impact of OTC concentration; (c) degradation efficiency; and (d) linear fitting of dynamic constants. Experimental conditions: [CaO2] = 0.05 g/L, [O3] = 0.75 g/h, [ HCO 3 ] = 2.25 mmol/L.
Figure 2. Impact of pH value: (a) degradation efficiency; (b) linear fitting of dynamic constants, impact of OTC concentration; (c) degradation efficiency; and (d) linear fitting of dynamic constants. Experimental conditions: [CaO2] = 0.05 g/L, [O3] = 0.75 g/h, [ HCO 3 ] = 2.25 mmol/L.
Molecules 29 00659 g002
Figure 3. Effect of scavenger: (a) methanol; (b) DABCO; (c) P-benzoquinone; and (d) indole. Experimental conditions: [CaO2] = 0.05 g/L, [O3] = 0.75 g/h, [ HCO 3 ] = 2.25 mmol/L, and [OTC] = 80 mmol/L.
Figure 3. Effect of scavenger: (a) methanol; (b) DABCO; (c) P-benzoquinone; and (d) indole. Experimental conditions: [CaO2] = 0.05 g/L, [O3] = 0.75 g/h, [ HCO 3 ] = 2.25 mmol/L, and [OTC] = 80 mmol/L.
Molecules 29 00659 g003
Figure 4. Formation of ·OH (a); ESR spectra of (b) ·OH and (c) 1O2.
Figure 4. Formation of ·OH (a); ESR spectra of (b) ·OH and (c) 1O2.
Molecules 29 00659 g004
Figure 5. 3D EEMF spectra of OTC samples under different treatment times: (a) 0 min; (b) 10 min; (c) 20 min; abd (d) 30 min.
Figure 5. 3D EEMF spectra of OTC samples under different treatment times: (a) 0 min; (b) 10 min; (c) 20 min; abd (d) 30 min.
Molecules 29 00659 g005
Figure 6. Degradation efficiency of (a) TOC and (b) COD during OTC degradation.
Figure 6. Degradation efficiency of (a) TOC and (b) COD during OTC degradation.
Molecules 29 00659 g006
Figure 7. (a) Fukui function of OTC mapped electron density isosurface: f; f+; and f0, (b) pathways for OTC degradation and (c) energy barrier diagram. (Gray atoms represent C, white atoms represent H, red atoms represent N, and blue atoms represent O).
Figure 7. (a) Fukui function of OTC mapped electron density isosurface: f; f+; and f0, (b) pathways for OTC degradation and (c) energy barrier diagram. (Gray atoms represent C, white atoms represent H, red atoms represent N, and blue atoms represent O).
Molecules 29 00659 g007
Figure 8. Toxicity of OTC degradation intermediates. (A: C22H24N2O10; B: C15H13NO7; C: C4H7NO4; D: C3H7NO4; E: C3H7NO3; F: C6H10O2; G: C6H12O).
Figure 8. Toxicity of OTC degradation intermediates. (A: C22H24N2O10; B: C15H13NO7; C: C4H7NO4; D: C3H7NO4; E: C3H7NO3; F: C6H10O2; G: C6H12O).
Molecules 29 00659 g008
Figure 9. Degradation of (a) BPA and (b) TC; (c) actual wastewater treatment situation.
Figure 9. Degradation of (a) BPA and (b) TC; (c) actual wastewater treatment situation.
Molecules 29 00659 g009aMolecules 29 00659 g009b
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, Z.; Xiang, L.; Pan, S.; Zhu, D.; Li, S.; Guo, H. The Degradation of Aqueous Oxytetracycline by an O3/CaO2 System in the Presence of HCO3: Performance, Mechanism, Degradation Pathways, and Toxicity Evaluation. Molecules 2024, 29, 659. https://doi.org/10.3390/molecules29030659

AMA Style

Li Z, Xiang L, Pan S, Zhu D, Li S, Guo H. The Degradation of Aqueous Oxytetracycline by an O3/CaO2 System in the Presence of HCO3: Performance, Mechanism, Degradation Pathways, and Toxicity Evaluation. Molecules. 2024; 29(3):659. https://doi.org/10.3390/molecules29030659

Chicago/Turabian Style

Li, Zedian, Liangrui Xiang, Shijia Pan, Dahai Zhu, Shen Li, and He Guo. 2024. "The Degradation of Aqueous Oxytetracycline by an O3/CaO2 System in the Presence of HCO3: Performance, Mechanism, Degradation Pathways, and Toxicity Evaluation" Molecules 29, no. 3: 659. https://doi.org/10.3390/molecules29030659

Article Metrics

Back to TopTop