Next Article in Journal
Metabolomics Provides Novel Insights into the Potential Toxicity Associated with Heated Tobacco Products, Electronic Cigarettes, and Tobacco Cigarettes on Human Bronchial Epithelial BEAS-2B Cells
Next Article in Special Issue
Sequential Treatment by Ozonation and Biodegradation of Pulp and Paper Industry Wastewater to Eliminate Organic Contaminants
Previous Article in Journal
In Vitro Toxicity Screening of Fifty Complex Mixtures in HepG2 Cells
Previous Article in Special Issue
Z-Type Heterojunction MnO2@g-C3N4 Photocatalyst-Activated Peroxymonosulfate for the Removal of Tetracycline Hydrochloride in Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study on the Direct and Indirect Photolysis of Antibacterial Florfenicol in Water Using DFT/TDDFT Method and Comparison of Its Reactivity with Hydroxyl Radical under the Effect of Metal Ions

Collaborative Innovation Center of Atmospheric Environment and Equipment Technology, Jiangsu Key Laboratory of Atmospheric Environment Monitoring and Pollution Control, School of Environmental Science and Engineering, Nanjing University of Information Science and Technology, Nanjing 210044, China
*
Author to whom correspondence should be addressed.
Toxics 2024, 12(2), 127; https://doi.org/10.3390/toxics12020127
Submission received: 7 January 2024 / Revised: 25 January 2024 / Accepted: 1 February 2024 / Published: 3 February 2024

Abstract

:
Florfenicol (FLO) is a widely used antibacterial drug, which is often detected in the environment. In this paper, the photolysis mechanism of FLO in water was investigated using density functional theory (DFT) and time-dependent density functional theory (TDDFT). The focus of the study is to elucidate the direct photolysis mechanism of FLO in the water environment and the indirect photolysis of free radicals (·OH, ·NO3, and ·SO4) as active species. The effect of metal ions Ca2+/Mg2+/Zn2+ on the indirect photolysis was also investigated. The results show that the direct photolysis of FLO involves C–C/C–N/C–S bond cleavage, the C5–S7 bond cleavage is most likely to occur, and the C17–C18 cleavage reaction is not easy to occur during the direct photodegradation of FLO. The indirect photolysis of FLO is more likely to occur in the environment than direct photolysis. The main indirect photolysis involves OH-addition, NO3-addition, and SO4-addition on benzene ring. The order of difficulty in the indirect photolysis with ·OH is C2 > C3 > C4 > C5 > C6 > C1, Ca2+ can promote the indirect photolysis with ·OH, and Mg2+/Zn2+ has a dual effect on the indirect photolysis with ·OH. In other words, Mg2+ and Zn2+ can inhibit or promote the indirect photolysis with ·OH. These studies provide important information for theoretical research on the environmental behavior and degradation mechanism of drug molecules.

1. Introduction

Florfenicol (FLO) is a broad-spectrum chloramphenicol antibiotic widely used in the prevention and treatment of bacterial diseases, which can inhibit or eliminate the growth of microorganisms [1,2,3,4,5]. During aquaculture, FLO, as a pharmaceutical feed, is usually leached directly or released into the aquatic environment with feces [6,7]. The average and maximum concentration of FLO in the water around the farm is 0.42 and 2.84 μg/L, and the maximum concentration in the Huangpu River is 46.6 ng L−1 [8,9]. Its residues are likely to lead to the development and spread of drug-resistant bacteria/resistance genes, which in turn limits the treatment options for human infections [10,11].
Since FLO exists in water for a long time and may pose a threat to the aquatic environment and food safety, there is an urgent need to investigate its chemical behavior in water to evaluate its risks [12,13]. The degradation of FLO in water mainly includes hydrolysis, biodegradation, and photolysis [14,15,16]. Photolysis can be divided into direct photolysis and indirect photolysis. Direct photolysis means that substances directly absorb light energy for decomposition reactions. During the indirect photolysis process, FLO can react with ·OH, O3, 1O2, and other reactive oxygen species in the water environment [17]. Direct and indirect photolysis play an indispensable role in the environmental behavior of drugs [14,18].
In recent years, some studies have focused on the degradation behavior of FLO under UV/H2O2, UV/Fe (II), and UV/Na2S2O3 conditions [18,19]. Xu et al. [20] studied the degradation behavior of FLO and its metabolites under different soil and natural conditions based on a simple method. In the past, photocatalysis has been shown to be an efficient treatment for the removal of aquaculture antibiotics, and it is necessary to consider changes in toxicity due to FLO decomposition [21,22]. Liu et al. [18] concluded that the toxicity of final products of FLO by UV irradiation was lower than FLO. UV irradiation-activated persulfate process can effectively control the toxicity of FLO degradation [19]. Meanwhile, photolysis is dramatically affected by environmental factors such as the type of light, pH, or antibiotic initial concentration [21,23,24]. Jiang et al. [25] considered that the superior stability and low concentration of C–F bonds in FLO hindered its degradation, and the degradation products may be as active/toxic as FLO. GE et al. [24] studied the effects of Cl, humic acid (HA), and other water components on the photodegradation of antibiotics; under UV–visible irradiation, Cl can promote the formation of 1O2 and accelerate the photodegradation of phenols, while under simulated solar irradiation, its photodegradation is not related to Cl; HA inhibits photolysis under UV–visible irradiation, but it can undergo photosensitive degradation under simulated solar irradiation. Laboratory experiments are time-consuming, laborious, and costly, and it is difficult to decipher all the pathways of organic pollutant degradation in a short time. More importantly, theoretical calculations can find the intermediate products of the reaction process, which is difficult to achieve with laboratory experiments. As a result, there is a lack of computational research on the water environment behavior of FLO, especially the photodegradation behavior in water environment, which has guiding significance and verification function for the laboratory study.
By the means of quantum chemical calculation, this study intends to use density functional theory (DFT) and time density functional theory (TDDFT) to explore the photochemical transformation process of FLO in water environment [26,27,28,29,30]. The focus of the study is to elucidate the direct photolysis mechanism of FLO in water and the indirect photolysis of free radicals (·OH, ·NO3, ·SO4) as active species. In addition, the effects of different metal ions (Ca2+/Mg2+/Zn2+) on the indirect photodegradation of FLO and ·OH were also studied. The computational methods employed in this research provide valuable insights into the photodegradation pathways and reactive intermediates involved in the photolysis of FLO in aqueous environments. Additionally, these findings contribute to a better understanding of the environmental fate and the potential remediation strategies of FLO contaminants.

2. Computational Methods

The calculations in this study were performed using the Gaussian 16 quantum chemistry software package [31,32]. The DFT/TDDFT has been extensively tested and validated against experimental data and has been shown to provide reliable predictions for the photolysis behavior of a variety of organic compounds [33,34,35,36]. Based on the B3LYP/6-311+G (d,p) level, DFT and TDDFT were used to optimize the stable structures of FLO in the ground state (S0) and the lowest excited triplet state (T1), respectively [37,38,39]. The intrinsic reaction coordinate (IRC) was used to analyze and verify the specified reactants (R) and products (P) for each transition state (TS) linkage [40]. In all calculations, the solvent effect in water is taken into account, and the integral equation in the form of the polarized continuum model (IEFPCM) is used [39]. At the B3LYP/6-311+G(d,p) level, TDDFT theory is used to calculate the electronic absorption spectra of monomer FLO and metal ion complexes FLO–Ca2+/FLO–Mg2+/FLO–Zn2+ in water [41,42]. The photodegradation mechanism of FLO in aqueous solution was analyzed by calculating the atomic charge and electron spin density based on natural bond orbital (NBO) at the B3LYP/6-311+G (d,p) level of theory. It mainly includes direct photodegradation (C–C bond cleavage, C–N bond cleavage, C–S bond cleavage), indirect photodegradation with free radicals (·OH, ·NO3, ·SO4) in water, and indirect photodegradation with ·OH under the effect of metal Ca2+/Mg2+/Zn2+.
The structures of Rs (reactants), TSs (transition states), and Ps (products) during the direct and indirect photolysis were optimized at the B3LYP/6-311+G(d,p) level. As T1 state has been found to be a precursor of photochemical reactions that exists for a long time, the calculation of photodegradation processes in this paper is carried out in T1 state [16]. Through the calculation and analysis of TS vibration, it is ensured that there is only one imaginary frequency. All the calculated activation energy (Ea) and enthalpy changes (ΔH) of the reaction are modified by zero–point energy.

3. Results and Discussion

3.1. Optimized Geometry of FLO

Figure 1a shows the optimized geometry of FLO. The calculation results indicate that although the benzene ring is symmetrical, the bond lengths are not the same. For example, the C2–C3 bond (1.466 Å) is longer than the C2–C1 bond (1.434 Å), and the C5–C4/C4–C3 bond (1.470 Å / 1.356 Å) is shorter than the C5–C6/C6–C1 bond (1.473 Å / 1.360 Å). Figure 1b shows the electronic absorption spectra of FLO. The calculated maximum absorption wavelength of FLO is 229 nm. Experimental results show that the maximum absorption wavelength of FLO is 224 nm [11]. This indicates that the theoretical calculation results are consistent with the experimental results. Due to absorption by the atmosphere, the shortest wavelength of solar radiation reaching the Earth’s surface is 290 nm [43]. In natural water, FLO can partially undergo direct photodegradation to a certain extent.

3.2. Direct Photolysis of FLO in Water

The direct photodegradation pathway of FLO in water with four reaction sites is shown in Figure S1. Figure S2 shows the optimized geometries of Rs, TSs, and Ps for the four reaction sites. The distance between the two sites of the broken bonds in the TSs for the four reaction sites is 1.799–2.100 Å. The ΔH values show that the cleavage of C11–C13/C5–S7 bond (Path 1/Path 4) is exothermic, and the cleavage of C13–N16/C17–C18 bond (Path 2/Path 3) is endothermic. The C11–C13 bond cleavage (Path 1) released the most heat (ΔH = −36.2 kcal/mol), indicating that the products P1a and P1(b) were the most stable.
Pathways 1, 2, 3, and 4 involve C11–C13, C13–N16, C17–C18, and C5–S7 bond cleavage (Path 1, Path 2, Path 3, and Path 4). The energy barrier diagram of the direct photolysis of FLO is shown in Figure 2. The Ea values of C11–C13/C13–N16/C5–S7 bond cleavage (Path 1/Path 2/Path 4) of FLO in T1 state are 10.0 kcal/mol, 35.0 kcal/mol, and 2.0 kcal/mol, respectively, which were obviously lower than the cleavage of C17–C18 bond (Ea = 94.8 kcal/mol) (Path 3). The relatively high Ea value (94.8 kcal/mol) indicates that the C17–C18 cleavage reaction (Path 3) does not easily occur during the direct photodegradation of FLO. The Ea value (2.0 kcal/mol) for the C5–S7 bond cleavage (Path 4) is the lowest among all the four paths, indicating that C5–S7 bond cleavage (Path 4) is most likely to occur in direct photodegradation, and this reaction produces P4a and P4b products.

3.3. Indirect Photolysis of FLO in Water

3.3.1. Indirect Photolysis Mechanism of FLO and ·OH

Six reaction pathways of FLO with ·OH in water were predicted, and the reaction type of these pathways was ·OH addition reaction (Figure 3). By calculation, the reactions at the C1, C2, C3, C4, C5, and C6 sites on the benzene ring were predicted to be likely to occur. Figure S3 shows the optimized geometries of Rs, TSs, and Ps at six reaction sites. The distance between ·OH and the six reaction sites in the TSs of FLO is 2.011–3.006 Å. Figure 3 shows that all the reaction pathways (Path C1, Path C2, Path C3, Path C4, Path C5, and Path C6) are exothermic, and the ΔH values range from −2.7 kcal/mol to −41.3 kcal/mol. The addition reaction (Path C5) at the C5 site is the most exothermic (ΔH = −41.3 kcal/mol), indicating that the generated products C5_Pa and C5_Pb are the most stable.
OH is predicted to attack six C sites on the benzene ring (Path C1, Path C2, Path C3, Path C4, Path C5, and Path C6) to generate C1_P, C2_P, C3_P, C4_P, C5_P, and C6_P. Zhang et al. [17] also detected that phenolic products were formed after the electrophilic attack of ·OH at the C site of benzene ring in FLO photodegradation experiments. The reaction of ·OH attacking at the C2 site (Path C2) releases the least heat (ΔH = −2.7 kcal/mol). Ea values range from 2.7 kcal/mol to 7.8 kcal/mol, and lower Ea values indicated that ·OH promotes the photolysis of FLO, which is consistent with the results of Li et al. [44]. The Ea value for ·OH (Path C1) attacking at the C1 reaction site is the lowest (2.7 kcal/mol), indicating that this reaction is the easiest to occur and most conducive to the formation of C1_P. In addition, since the Ea value (7.8 kcal/mol) for ·OH attacking at C2 (Path C2) is the maximum value of all the reactions of ·OH attacking at FLO, Path C2 is not conducive to occur. A similar reaction was also shown at the C3 site. The Ea value for ·OH attacking at the C3 site (Path C6) is 7.3 kcal/mol. The data with little difference indicate that C2_P and C3_P are likely to be generated simultaneously

3.3.2. Indirect Photolysis Mechanism of FLO and ·NO3

Three possible pathways (Path C1N, Path C4N, and Path C6N) of the indirect photodegradation of FLO by ·NO3 in water are shown in Figure 4, and the reaction type of these pathways is an addition reaction. There are three possible reaction sites on the benzene ring: C1, C4, and C6. Figure S4 shows the geometric optimization of Rs, TSs, and Ps for the three reaction sites. The distance between ·NO3 and the three reaction sites in the TSs is 1.711–1.866 Å. Figure 4 shows that the three reaction pathways (Path C1N, Path C4N, and Path C6N) are endothermic, with ΔH values ranging from 4.4 kcal/mol to 4.8 kcal/mol.
The relatively low Ea value (5.5 kcal/mol) indicates that ·OH is most likely to react at the C1 site (Path C1N) and is most conducive to the formation of C1N_Pa. Similar reactions were also shown at the C4 and C6 sites. The Ea values for adding ·OH at the C4 site (Path C4N) and C6 site (Path C6N) are 6.1 kcal/mol and 5.9 kcal/mol, respectively. The data with little difference show that sites C1, C4, and C6 are beneficial to react with ·OH.

3.3.3. Indirect Photolysis Mechanism of FLO and ·SO4

Four possible pathways (Path C1S, Path C3S, Path C4S, and Path C6S) of the indirect photodegradation of FLO with ·SO4 in water are shown in Figure 5, and the reaction type is an addition reaction. Figure S5 shows the optimized geometries of Rs, TSs, and Ps for the four reaction sites. The distance between ·SO4 and the four reaction sites in the TSs is 1.910–1.964 Å. All the pathways (Path C1S, Path C3S, Path C4S, and Path C6S) are predicted to be endothermic, with ΔH ranging from 4.5 kcal/mol to 8.2 kcal/mol (Figure 5).
The lowest Ea value (12.7 kcal/mol) for Path C6S among all the four pathways indicates that ·OH is most likely to react at the C6 site and is most conducive to the formation of C6S_Pa. The Ea value for adding ·SO4 to the C1 site was 15.9 kcal/mol (Path C1S), which is the maximum value of all the four pathways. This indicates that the reaction of ·SO4 attacking at the C1 site (Path C1S) is not conducive to proceed.

3.3.4. Optimized Geometries of Complexes FLO–Ca2+/FLO–Mg2+/FLO–Zn2+

Figure S6 shows the optimized geometries of complexes FLO–Ca2+/FLO–Mg2+/FLO–Zn2+. The calculation results show that the bond lengths of the complexes are different from those of the monomer FLO. For example, the bond lengths of S7–O9/ S7–C5/C2–C11 in complex FLO–Mg2+ are longer than those of FLO, conversely, the bond lengths of C11–C13 in complex FLO–Mg2+ are shorter than those of FLO (Figure 1a and Figure S9).
Figure 1b shows the electronic absorption spectra of complexes FLO–Ca2+/FLO–Mg2+/FLO–Zn2+ in water at the TDDFT/B3LYP/6-311+G(d,p) level of theory. The bond length of Ca–O8 in the complex FLO–Ca2+ is the longest (2.329 Å) and the maximum absorption wavelength is the smallest (189 nm) among all the three metal ions complexes. The bond length of Zn–O8 in the complex FLO–Zn2+ is the shortest (1.926 Å), and the maximum absorption wavelength is the largest (238 nm).
Compared with the monomer, there is a blue shift in the maximum electron absorbance peak position of complex FLO–Ca2+, while there is a red shift in those of the complexes FLO–Mg2+/FLO–Zn2+. Therefore, the metal ions Ca2+/Mg2+/Zn2+ have an effect on the structure and UV absorption spectrum of FLO and may also have an effect on the photodegradation of FLO.

3.3.5. Indirect Photolysis of Complexes FLO–Ca2+/FLO–Mg2+/FLO–Zn2+ with ·OH

Mechanism of Indirect Photolysis of Complex FLO–Ca2+ with ·OH in Water

Figure S7 shows the optimized geometries of complex FLO–Ca2+. Results show that the complex FLO–Ca2+ has four possible geometries. By calculating the minimum value of the potential energy surface, the most stable geometric structure (FLO–Ca2+–OPT3) with the lowest single-point energy is obtained among the four structures (Figure S8). All of the following calculations are based on the FLO–Ca2+–OPT3 structure.
Two possible pathways (Path C3Ca and Path C4Ca) of the indirect photolysis of complex FLO–Ca2+ with ·OH are shown in Figure 6, and the reaction type is an addition reaction. There are two possible reaction sites on the benzene ring: C3 and C4. Figure S7 shows the optimized geometries of Rs, TSs, and Ps for the two reaction sites. The distance between ·OH and the two reaction sites in the TSs is 2.028 and 2.033 Å, and the difference is very small. Both the two pathways (Path C3Ca and Path C4Ca) shown in Figure S7 are exothermic, with ΔH values ranging from −8.8 to −9.3 kcal/mol.
The lower Ea value (3.8 kcal/mol) indicates that ·OH is more likely to react at the C3 site (Path C3Ca) than that at the C4 site. The Ea value for adding ·OH to the C4 site (Path C4Ca) is 4.3 kcal/mol. For the complex FLO–Ca2+, the calculated Ea values for the two pathways by ·OH (Path C3Ca and Path C4Ca) are lower than that for the indirect photolysis of monomer FLO. This indicates that Ca2+ can promote the ·OH addition reaction at C3 and C4 sites. In addition, as shown in Figure 3, ·OH is more prone to attack the C4 site than the C3 site for the indirect photolysis of monomer FLO. However, in the presence of Ca2+, the Ea value for adding ·OH to the C3 site (Path C3Ca, Ea = 3.8 kcal/mol) is lower than that for the C4 site (Path C4Ca, Ea = 4.3 kcal/mol). This indicates that Ca2+ can change the main indirect photolysis pathway of FLO.

Indirect Photodegradation Mechanism of Complex FLO–Mg2+ and ·OH in Water

The optimized geometries of complex FLO–Mg2+ are shown in Figure S9. The results show that complex FLO–Mg2+ has five possible geometries. The most stable geometric structure (FLO–Mg2+–OPT4 structure) with the lowest single-point energy is obtained among the five structures (Figure S9). All calculations are based on the FLO–Mg2+–OPT4 structure.
Four possible pathways (Path C3Mg, Path C4Mg, Path C5Mg, and Path C6Mg) of the indirect photodegradation of complex FLO–Mg2+ with ·OH are shown in Figure 7, and the reaction type of these pathways is an addition reaction. Four reaction sites (C3, C4, C5, and C6) on the benzene ring were predicted to undergo reactions. Figure S10 shows the optimized geometries of Rs, TSs, and Ps for these four reaction sites. The distance between ·OH and the four reaction sites in the TSs is 1.998–2.049 Å. Figure 7 shows that the ΔH values for all the pathways (Path C3Mg, Path C4Mg, Path C5Mg, and Path C6Mg) range from −6.4 kcal/mol to −36.4 kcal/mol, and the four pathways are exothermic. The ΔH value (−36.4 kcal/mol) for ·OH attacking at the C5 site is the lowest among the four pathways, indicating that the products C5Mg_Pa and C5Mg_Pb are the most stable.
The Ea value (2.9 kcal/mol) for Path C4Mg of all the four pathways is the lowest and indicates that ·OH is most likely to react at the C4 site. A similar reaction was also obtained at the C6 site. The Ea value for adding ·OH at the C6 site is 4.3 kcal/mol. In addition, the Ea value (6.8 kcal/mol) for ·OH attacking at the C3 site is the highest of all the four pathways, so Path C3Mg is the most difficult for undergoing photolysis reactions with ·OH. For C3 and C4 sites, the calculated Ea values of the indirect photolysis of the complex FLO–Mg2+ with ·OH (Path C3Mg and Path C4Mg) are lower than those of monomer FLO with ·OH(Path C3 and Path C4). In addition, Ea values for the Path C5Mg and Path C6Mg of complex FLO–Mg2+ are higher than those of the reaction of monomer FLO with ·OH at C5 and C6 sites(Path C5 and Path C6). This indicates that Mg2+ has a dual role in the photolysis of FLO. Mg2+ can promote ·OH attacking at the C3 and C4 sites (Path C3Mg and Path C4Mg) and inhibit ·OH attacking at the C5 and C6 sites (Path C5Mg and Path C6Mg). In addition, Mg2+ can also change the main photolysis pathway of FLO with ·OH. In the presence of Mg2+, the order of difficulty in the indirect photolysis with ·OH is C3 > C5 > C6 > C4, and for monomer FLO, the order is C3 > C4 > C5 > C6.

Mechanism of Indirect Photodegradation of Complex FLO–Zn2+ with ·OH in Water

The optimized geometries of complex FLO–Zn2+ are shown in Figure S11. The results show that the complex FLO–Zn2+ has four possible geometries, and the most stable geometric structure with the lowest single-point energy is FLO–Zn2+–OPT1 (Figure S11). Thus, all calculations are based on the FLO–Zn2+–OPT1 structure.
Four possible pathways (Path C1Zn, Path C3Zn, Path C4Zn, and Path C6Zn) of the indirect photodegradation of the complex FLO–Zn2+ with ·OH are shown in Figure 8, and the reaction type is an addition reaction. Figure S12 shows the optimized geometries of Rs, TSs, and Ps for the four reaction sites (C1, C3, C4, and C6). The distance between ·OH and the four reaction sites is 2.013–2.036 Å. Figure 8 shows that ΔH values for all the four pathways (Path C1Zn, Path C3Zn, Path C4Zn, and Path C6Zn) range from −9.4 kcal/mol to −11.6 kcal/mol, and these pathways are exothermic. The ΔH value (−11.6 kcal/mol) for ·OH attacking at the C6 site is the lowest among all the four pathways, indicating that product C6Zn_P is the most stable.
The Ea values for ·OH attacking at the C1, C3, and C6 sites are all 3.1 kcal/mol, which are lower than that for the C4 cite (4.5 kcal/mol). This indicates that ·OH is more likely to react at the C1, C3, and C6 sites (Path C1Zn, Path C3Zn, and Path C6Zn). In addition, the calculated Ea values for Path C3Zn, Path C4Zn, and Path C6Zn of the complex FLO–Zn2+ are lower than those for the indirect photolysis of monomer FLO with ·OH, and the Ea value for Path C1Zn of complex FLO–Zn2+ is higher than that for the indirect photolysis of monomer FLO with ·OH. This indicates that Zn2+ has a dual role in the indirect photolysis of FLO with ·OH, i.e., Zn2+ can promote ·OH attacking at the C3, C4, and C6 sites and inhibit ·OH attacking at the C1 site. In addition, Zn2+ can also change the main photodegradation pathway of FLO with ·OH, i.e., in the presence of Zn2+, the order of difficulty of ·OH attacking at the different reaction sites in the indirect photodegradation of FLO with ·OH is C4 > C6 = C1 = C3, which is different from that of monomer FLO (C3 > C4 > C6 > C1).

4. Conclusions

This study provides a basis for further understanding the photodegradation mechanism of FLO in water. These mechanisms involve direct photodegradation, indirect photodegradation with free radicals (·OH, ·NO3, and ·SO4), and the effect of metal ions (Ca2+/Mg2+/Zn2+) on indirect photodegradation with ·OH. The calculation results show that the direct photodegradation of FLO involves three reaction types (C–C/C–N/C–S bond cleavage). The relatively high Ea value (94.8 kcal/mol) indicates that C17–C18 cleavage cannot easily occur during the direct photodegradation of FLO. The lower Ea value (2.0 kcal/mol) indicates that C5–S7 cleavage is the main pathway of the direct photodegradation process. The indirect photodegradation of FLO by free radicals is an addition reaction. The ΔH (−41.3 kcal/mol) for adding ·OH at the C5 site is higher than that for other sites on the benzene ring, indicating that the products such as 2,2-dichloro-N-(3-fluoro-1-hydroxy-1-(4-hydroxyphenyl) propan-2-yl) acetamide are the most stable. The indirect photodegradation of FLO by ·NO3 and ·SO4 is endothermic. The NO3-addition reaction is the most likely to occur at the C1 site, while the ·SO4-addition reaction is the least likely to occur at the C1 site. The metal ion Ca2+ can promote the indirect photodegradation with ·OH, and the metal ion Mg2+/ Zn2+ has a dual effect on the indirect photodegradation with ·OH. Mg2+ can not only promote the ·OH addition reaction at C3 and C4 sites but also inhibit the ·OH addition reaction at C5 and C6 sites. Zn2+ can not only promote the ·OH addition reaction at C3, C4, and C6 sites but also inhibit the ·OH addition reaction at C1 site. In addition, in the presence of metal ion Mg2+, OH-addition is the most likely to occur at the C4 site, but it is the least likely to occur at the C3 site. These results indicate that the quantum chemical calculation method can be used to study the mechanism of the water-based photochemical transformation of antibacterial drugs and to guide their safe use and risk avoidance.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/toxics12020127/s1, Figure S1: Direct photodegradation pathways of FLO, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol). The numbers in parentheses are NBO charge and electron spin density (NBO charge, electron spin density); Figure S2: Optimized geometries of reactants, transition states, and products in direct photolysis pathways (Path 1, Path 2, Path 3, and Path 4) of FLO, along with selected bond lengths (Å); Figure S3: Optimized geometries of reactants, transition states, and products in indirect photolysis pathways (Path C1, Path C2, Path C3, and Path C4) of FLO with ·OH, along with selected bond lengths (Å); Figure S4: Optimized geometries of reactants, transition states, and products in indirect photolysis pathways (Path C1N, Path C4N, and Path C6N) of FLO with ·NO3, along with selected bond lengths (Å); Figure S5: Optimized geometries of reactants, transition states, and products in indirect photolysis pathways (Path C1S, Path C3S, Path C4S, and Path C6S) of FLO with ·SO4-, along with selected bond lengths (Å); Figure S5: Optimized geometries of complexes FLO–Ca2+/FLO–Mg2+/FLO–Zn2+, along with selected bond lengths (Å); Figure S7: Four optimized geometries (OPT1, OPT2, OPT3, and OPT4) of complex FLO–Ca2+ along with selected bond length (Å). The energies of geometries are relative to that of the most stable geometry FLO–Ca2+–OPT3; Figure S8: Optimized geometries of reactants, transition states, and products in indirect photolysis pathways (Path C3Ca and Path C4Ca) of complex FLO–Ca2+ with ·OH, along with selected bond lengths (Å); Figure S9: Five optimized geometries (OPT1, OPT2, OPT3, OPT4, and OPT5) of complex FLO–Mg2+ along with selected bond length (Å). The energies of geometries are relative to that of the most stable geometry FLO–Mg2+–OPT4; Figure S10: Optimized geometries of reactants, transition states, and products in indirect photolysis pathways (Path C3Mg, Path C4Mg, Path C5Mg, and Path C6Mg) of complex FLO–Mg2+ with ·OH, along with selected bond lengths (Å); Figure S11: Four optimized geometries (OPT1, OPT2, OPT3, and OPT4) of complex FLO–Zn2+ along with selected bond length (Å). The energies of geometries are relative to that of the most stable geometry FLO–Zn2+–OPT1; Figure S12: Optimized geometries of reactants, transition states, and products in indirect photolysis pathways (Path C1Zn, Path C3Zn, Path C4Zn, and Path C6Zn) of complex FLO–Zn2+ with ·OH, along with selected bond lengths (Å).

Author Contributions

Y.K. performed the theoretical calculations and wrote the manuscript; Y.L. performed the data analyses and manuscript preparation; S.W. contributed to the conception of the study and wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (41601519) and the Jiangsu Students’ Platform for innovation and entrepreneurship training program (202210300099Y). Thank the NUIST Students’ Platform for Innovation and Entrepreneurship Training Program for financial support (XJDC202210300157).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request. The dataset used and/or analyzed in this study is available from the corresponding author on reasonable request.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Schunicht, O.C.; Booker, C.W.; Guichon, P.T.; Jim, G.K.; Wildman, B.K.; Pittman, T.J.; Perrett, T. An evaluation of the relative efficacy of tulathromycin for the treatment of undifferentiated fever in feedlot calves in Nebraska. Can. Vet. J.-Rev. Vet. Can. 2007, 48, 600–606. [Google Scholar]
  2. Leng, N.N.; Ju, M.J.; Jiang, Y.T.; Guan, D.; Liu, J.H.; Chen, W.; Algharib, S.A.; Dawood, A.; Luo, W.H. The therapeutic effect of florfenicol-loaded carboxymethyl chitosan-gelatin shell nanogels against Escherichia coli infection in mice Running title: Therapeutic effect of florfenicol shell nanogels. J. Mol. Struct. 2022, 1269, 7. [Google Scholar] [CrossRef]
  3. Li, Z.; Yang, Y.J.; Qin, Z.; Li, S.H.; Bai, L.X.; Li, J.Y.; Liu, X.W. Florfenicol-Polyarginine Conjugates Exhibit Promising Antibacterial Activity Against Resistant Strains. Front. Chem. 2022, 10, 13. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, J.H.; Ju, M.J.; Wu, Y.F.; Leng, N.N.; Algharib, S.A.; Luo, W.H. Antibacterial activity of florfenicol composite nanogels against Staphylococcus aureus small colony variants. J. Vet. Sci. 2022, 23, 13. [Google Scholar] [CrossRef]
  5. Park, B.-K.; Lim, J.-H.; Kim, M.-S.; Hwang, Y.-H.; Yun, H.-I. Pharmacokinetics of florfenicol and its metabolite, florfenicol amine, in dogs. Res. Vet. Sci. 2008, 84, 85–89. [Google Scholar] [CrossRef]
  6. Barreto, F.M.; da Silva, M.R.; Braga, P.A.C.; Bragotto, A.P.A.; Hisano, H.; Reyes, F.G.R. Evaluation of the leaching of florfenicol from coated medicated fish feed into water. Environ. Pollut. 2018, 242, 1245–1252. [Google Scholar] [CrossRef] [PubMed]
  7. Pouliquen, H.; Delpépée, R.; Larhantec-Verdier, M.; Morvan, M.L.; Le Bris, H. Comparative hydrolysis and photolysis of four antibacterial agents (oxytetracycline oxolinic acid, flumequine and florfenicol) in deionised water, freshwater and seawater under ablotic conditions. Aquaculture 2007, 262, 23–28. [Google Scholar] [CrossRef]
  8. Bu, Q.W.; Wang, B.; Huang, J.; Deng, S.B.; Yu, G. Pharmaceuticals and personal care products in the aquatic environment in China: A review. J. Hazard. Mater. 2013, 262, 189–211. [Google Scholar] [CrossRef]
  9. Wei, R.C.; Ge, F.; Chen, M.; Wang, R. Occurrence of Ciprofloxacin, Enrofloxacin, and Florfenicol in Animal Wastewater and Water Resources. J. Environ. Qual. 2012, 41, 1481–1486. [Google Scholar] [CrossRef]
  10. Martinez, J.L. Environmental pollution by antibiotics and by antibiotic resistance determinants. Environ. Pollut. 2009, 157, 2893–2902. [Google Scholar] [CrossRef]
  11. Martins, Y.A.; De Oliveira, C. Development and Validation of an UV Spectrophotometric Method for Determination of Thiamphenicol in Dosage Form. J. Appl. Spectrosc. 2019, 86, 629–635. [Google Scholar] [CrossRef]
  12. Sun, Z.J.; Zhang, L.W.; Dong, D.M.; Guo, Z.Y. Optimizing the multimedia fate model for characterizing environmental risks of florfenicol in seasonally ice-covered reservoirs. Environ. Pollut. 2023, 323, 10. [Google Scholar] [CrossRef]
  13. Yévenes, K.; Pokrant, E.; Pérez, F.; Riquelme, R.; Avello, C.; Maddaleno, A.; San Martín, B.; Cornejo, J. Assessment of Three Antimicrobial Residue Concentrations in Broiler Chicken Droppings as a Potential Risk Factor for Public Health and Environment. Int. J. Environ. Res. Public Health. 2018, 16, 24. [Google Scholar] [CrossRef] [PubMed]
  14. Ding, Y.C.; Liang, B.; Jiang, W.L.; Han, J.L.; Guadie, A.; Yun, H.; Cheng, H.Y.; Yang, R.J.; Liu, S.J.; Wang, A.J.; et al. Effect of preferential UV photolysis on the source control of antibiotic resistome during subsequent biological treatment systems. J. Hazard. Mater. 2021, 414, 10. [Google Scholar] [CrossRef] [PubMed]
  15. Song, C.F.; Wei, Y.L.; Qiu, Y.T.; Qi, Y.; Li, Y.; Kitamura, Y. Biodegradability and mechanism of florfenicol via Chlorella sp UTEX1602 and L38: Experimental study. Bioresour. Technol. 2019, 272, 529–534. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, H.M.; Wei, X.X.; Song, X.D.; Shah, S.; Chen, J.W.; Liu, J.H.; Hao, C.; Chen, Z.F. Photophysical and photochemical insights into the photodegradation of sulfapyridine in water: A joint experimental and theoretical study. Chemosphere 2018, 191, 1021–1027. [Google Scholar] [CrossRef] [PubMed]
  17. Zhang, Y.; Li, J.H.; Zhou, L.; Wang, G.Q.; Feng, Y.H.; Wang, Z.Y.; Yang, X. Aqueous photodegradation of antibiotic florfenicol: Kinetics and degradation pathway studies. Environ. Sci. Pollut. Res. 2016, 23, 6982–6989. [Google Scholar] [CrossRef] [PubMed]
  18. Liu, N.; Sijak, S.; Zheng, M.; Tang, L.; Xu, G.; Wu, M.H. Aquatic photolysis of florfenicol and thiamphenicol under direct UV irradiation, UV/H2O2 and UV/Fe(II) processes. Chem. Eng. J. 2015, 260, 826–834. [Google Scholar] [CrossRef]
  19. Gao, Y.Q.; Gao, N.Y.; Deng, Y.; Yin, D.Q.; Zhang, Y.S. Degradation of florfenicol in water by UV/Na2S2O8 process. Environ. Sci. Pollut. Res. 2015, 22, 8693–8701. [Google Scholar] [CrossRef]
  20. Xu, M.F.; Qian, M.R.; Zhang, H.; Ma, J.W.; Wang, J.M.; Wu, H.Z. Simultaneous determination of florfenicol with its metabolite based on modified quick, easy, cheap, effective, rugged, and safe sample pretreatment and evaluation of their degradation behavior in agricultural soils. J. Sep. Sci. 2015, 38, 211–217. [Google Scholar] [CrossRef]
  21. Silva, C.; Louros, V.; Silva, V.; Otero, M.; Lima, D. Antibiotics in Aquaculture Wastewater: Is It Feasible to Use a Photodegradation-Based Treatment for Their Removal? Toxics 2021, 9, 194. [Google Scholar] [CrossRef] [PubMed]
  22. Tian, J.; Li, B.B.; Qu, R.J.; Zhou, D.M.; Sun, C.; Wang, Z.Y.; Zhu, F. Influence of anions on ozonation of bisphenol AF: Kinetics, reaction pathways, and toxicity assessment. Chemosphere 2022, 286, 131864. [Google Scholar] [CrossRef] [PubMed]
  23. Malvaldi, M.; Bruzzone, S.; Chiappe, C.; Gusarov, S.; Kovalenko, A. Ab Initio Study of Ionic Liquids by KS-DFT/3D-RISM-KH Theory. J. Phys. Chem. B 2009, 113, 3536–3542. [Google Scholar] [CrossRef] [PubMed]
  24. Ge, L.; Chen, J.W.; Lin, J.; Cai, X.Y. Light-Source-Dependent Effects of Main Water Constituents on Photodegradation of Phenicol Antibiotics: Mechanism and Kinetics. Environ. Sci. Technol. 2009, 43, 3101–3107. [Google Scholar] [CrossRef]
  25. Jiang, W.L.; Ding, Y.C.; Haider, M.R.; Han, J.L.; Liang, B.; Xia, X.; Yang, L.M.; Wang, H.C.; Peng, Y.Z.; Wang, A.J. A novel TiO2/graphite felt photoanode assisted electro-Fenton catalytic membrane process for sequential degradation of antibiotic florfenicol and elimination of its antibacterial activity. Chem. Eng. J. 2020, 391, 12. [Google Scholar] [CrossRef]
  26. Liu, C.G.; Guan, X.H.; Su, Z.M. TDDFT study on the second-order nonlinear optical properties of a series of mono- and di-nuclear 60 fullerene complexes. Comput. Theor. Chem. 2011, 963, 98–103. [Google Scholar] [CrossRef]
  27. Minaev, B.F.; Baryshnikov, G.V.; Korop, A.A.; Minaeva, V.A.; Kaplunov, M.G. Theoretical investigation of the structure and electronic absorption spectrum of a complex zinc bis- 8-(3,5-difluorophenylsulfanylamino)quinolinate. Opt. Spectrosc. 2012, 113, 298–304. [Google Scholar] [CrossRef]
  28. Berland, K.; Londero, E.; Schröder, E.; Hyldgaard, P. Harris-type van der Waals density functional scheme. Phys. Rev. B 2013, 88, 10. [Google Scholar] [CrossRef]
  29. Wolny, J.A.; Paulsen, H.; Winkler, H.; Trautwein, A.X.; Tuchagues, J.P. DFT calculations as a tool to analyse quadrupole splittings of spin crossover Fe(II) complexes. Hyperfine Interact. 2005, 166, 495–498. [Google Scholar] [CrossRef]
  30. Shah, T.A.; Alam, U.; Alam, M.; Park, S.; Muneer, M. Single crystal X-ray structure, spectroscopic and DFT studies of Imidazo [2,1-b]thiazole: 2-(3-hydroxy-3-phenylimidazo[2,1-b]thiazol-2(3H)-ylidene)-1-phenylethanone. J. Mol. Struct. 2018, 1157, 638–653. [Google Scholar] [CrossRef]
  31. Benzon, K.B.; Mary, Y.S.; Varghese, H.T.; Panicker, C.Y.; Armakovic, S.; Armakovic, S.J.; Pradhan, K.; Nanda, A.K.; Van Alsenoy, C. Spectroscopic, DFT, molecular dynamics and molecular docking study of 1-butyl-2-(4-hydroxypheny1)-4,5-dimethyl-imidazole 3-oxide. J. Mol. Struct. 2017, 1134, 330–344. [Google Scholar] [CrossRef]
  32. Wilson, P.B.; Williams, I.H. Critical evaluation of anharmonic corrections to the equilibrium isotope effect for methyl cation transfer from vacuum to dielectric continuum. Mol. Phys. 2015, 113, 1704–1711. [Google Scholar] [CrossRef]
  33. Arrigoni, F.; Zampella, G.; Gioia, L.D.; Greco, C.; Bertini, L.J.I. The Photochemistry of Fe2(S2C3H6)(CO)6(-CO) and Its Oxidized Form, Two Simple [FeFe]-Hydrogenase CO-Inhibited Models. A DFT and TDDFT Investigation. Inorganics. 2021, 9, 16. [Google Scholar] [CrossRef]
  34. Ni, Z.; Chen, Z.; Zhang, X.; Yang, X.; Zhou, L. Photolysis of fungicide triadimefon: A combined experimental and theoretical investigation on homolytic C-O and C-N bonds dissociation mechanisms. J. Photochem. Photobiol. A Chem. 2023, 436, 114402. [Google Scholar] [CrossRef]
  35. Polska, K.; Zielonka, J.; Chomicz, L.; Czerwicka, M.; Stepnowski, P.; Guzow, K.; Wiczk, W.; Smużyńska, M.; Kasprzykowski, F.; Zylicz-Stachula, A.; et al. Unexpected photoproduct generated via the acetone-sensitized photolysis of 5-bromo-2’-deoxyuridine in a water/isopropanol solution: Experimental and computational studies. J. Phys. Chem. B 2010, 114, 16902–16907. [Google Scholar] [CrossRef]
  36. Sun, Y.; Li, M.; Hadizadeh, M.H.; Liu, L.; Xu, F. Theoretical insights into the degradation mechanisms, kinetics and eco-toxicity of oxcarbazepine initiated by OH radicals in aqueous environments. J. Environ. Sci. 2023, 129, 189–201. [Google Scholar] [CrossRef]
  37. Noh, S.H.; Kwak, D.H.; Seo, M.H.; Ohsaka, T.; Han, B. First principles study of oxygen reduction reaction mechanisms on N-doped graphene with a transition metal support. Electrochim. Acta 2014, 140, 225–231. [Google Scholar] [CrossRef]
  38. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  39. Tomasi, J.; Mennucci, B.; Cammi, R.J.C.R. Quantum mechanical continuum salvation models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef]
  40. Yamashita, K.; Yamabe, T.; Fukui, K. IRC approach to chemical dynamics: Toward mode-selective chemical reactions. Chem. Phys. Lett. 1981, 84, 123–126. [Google Scholar] [CrossRef]
  41. Zhao, G.J.; Han, K.L. Excited State Electronic Structures and Photochemistry of Heterocyclic Annulated Perylene (HAP) Materials Tuned by Heteroatoms: S, Se, N, O, C, Si, and B. J. Phys. Chem. A 2009, 113, 4788–4794. [Google Scholar] [CrossRef]
  42. Marques, M.A.L.; Gross, E.K.U. Time-Dependent Density Functional Theory. Annu. Rev. Phys. Chem. 2008, 55, 427–455. [Google Scholar] [CrossRef] [PubMed]
  43. Hölzle, E.; Hönigsmann, H. Ultraviolette Strahlung—Quellen, Spektren, Umwelteinflüsse: UV-radiation—Sources, Wavelength, Environment. JDDG J. Der Dtsch. Dermatol. Ges. 2010, 3, S3–S10. [Google Scholar] [CrossRef] [PubMed]
  44. Li, K.; Zhang, P.; Ge, L.K.; Ren, H.L.; Yu, C.Y.; Chen, X.Y.; Zhao, Y.F. Concentration-dependent photodegradation kinetics and hydroxyl-radical oxidation of phenicol antibiotics. Chemosphere 2014, 111, 278–282. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Optimized geometries of FLO along with selected bond lengths (Å) and (b) calculated electronic absorption spectra of monomer FLO and complexes FLO–Ca2+/FLO–Mg2+/FLO–Zn2+ with maximum absorption wavelength (nm).
Figure 1. (a) Optimized geometries of FLO along with selected bond lengths (Å) and (b) calculated electronic absorption spectra of monomer FLO and complexes FLO–Ca2+/FLO–Mg2+/FLO–Zn2+ with maximum absorption wavelength (nm).
Toxics 12 00127 g001
Figure 2. Reaction energy barrier diagram of the direct photolysis of FLO.
Figure 2. Reaction energy barrier diagram of the direct photolysis of FLO.
Toxics 12 00127 g002
Figure 3. Indirect photodegradation pathways of FLO with ·OH, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol).
Figure 3. Indirect photodegradation pathways of FLO with ·OH, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol).
Toxics 12 00127 g003
Figure 4. Indirect photodegradation pathways of FLO with ·NO3, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol).
Figure 4. Indirect photodegradation pathways of FLO with ·NO3, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol).
Toxics 12 00127 g004
Figure 5. Indirect photodegradation pathways of FLO with ·SO4, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol).
Figure 5. Indirect photodegradation pathways of FLO with ·SO4, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol).
Toxics 12 00127 g005
Figure 6. Indirect photodegradation pathways of complex FLO–Ca2+ with ·OH, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol).
Figure 6. Indirect photodegradation pathways of complex FLO–Ca2+ with ·OH, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol).
Toxics 12 00127 g006
Figure 7. Indirect photodegradation pathways of complex FLO–Mg2+ with ·OH, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol).
Figure 7. Indirect photodegradation pathways of complex FLO–Mg2+ with ·OH, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol).
Toxics 12 00127 g007
Figure 8. Indirect photodegradation pathways of complex FLO–Zn2+ with ·OH, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol).
Figure 8. Indirect photodegradation pathways of complex FLO–Zn2+ with ·OH, along with computed activation energies (Ea, kcal/mol) and enthalpy change (ΔH, kcal/mol).
Toxics 12 00127 g008
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

Kang, Y.; Lu, Y.; Wang, S. Study on the Direct and Indirect Photolysis of Antibacterial Florfenicol in Water Using DFT/TDDFT Method and Comparison of Its Reactivity with Hydroxyl Radical under the Effect of Metal Ions. Toxics 2024, 12, 127. https://doi.org/10.3390/toxics12020127

AMA Style

Kang Y, Lu Y, Wang S. Study on the Direct and Indirect Photolysis of Antibacterial Florfenicol in Water Using DFT/TDDFT Method and Comparison of Its Reactivity with Hydroxyl Radical under the Effect of Metal Ions. Toxics. 2024; 12(2):127. https://doi.org/10.3390/toxics12020127

Chicago/Turabian Style

Kang, Yue, Ying Lu, and Se Wang. 2024. "Study on the Direct and Indirect Photolysis of Antibacterial Florfenicol in Water Using DFT/TDDFT Method and Comparison of Its Reactivity with Hydroxyl Radical under the Effect of Metal Ions" Toxics 12, no. 2: 127. https://doi.org/10.3390/toxics12020127

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop