Next Article in Journal
Comparative Reactive Oxygen Species (ROS) Content among Various Flavored Disposable Vape Bars, including Cool (Iced) Flavored Bars
Next Article in Special Issue
Predicting the Bioconcentration Factor in Fish from Molecular Structures
Previous Article in Journal
Microbial Transglutaminase Is a Very Frequently Used Food Additive and Is a Potential Inducer of Autoimmune/Neurodegenerative Diseases
Previous Article in Special Issue
Potential Environmental Risk Characteristics of PCB Transformation Products in the Environmental Medium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Theoretical Calculation on the Reaction Mechanisms, Kinetics and Toxicity of Acetaminophen Degradation Initiated by Hydroxyl and Sulfate Radicals in the Aqueous Phase

1
Hubei Key Laboratory of Pollutant Analysis & Reuse Technology, College of Chemistry and Chemical Engineering, Hubei Normal University, Cihu Road 11, Huangshi 435002, China
2
Huangshi Key Laboratory of Photoelectric Technology and Materials, College of Physics and Electronic Science, Hubei Normal University, Huangshi 435002, China
3
College of Urban and Environmental Sciences, Hubei Normal University, Huangshi 435002, China
*
Author to whom correspondence should be addressed.
Toxics 2021, 9(10), 234; https://doi.org/10.3390/toxics9100234
Submission received: 18 August 2021 / Revised: 16 September 2021 / Accepted: 21 September 2021 / Published: 25 September 2021

Abstract

:
The •OH and SO4•− play a vital role on degrading pharmaceutical contaminants in water. In this paper, theoretical calculations have been used to discuss the degradation mechanisms, kinetics and ecotoxicity of acetaminophen (AAP) initiated by •OH and SO4•−. Two significant reaction mechanisms of radical adduct formation (RAF) and formal hydrogen atom transfer (FHAT) were investigated deeply. The results showed that the RAF takes precedence over FHAT in both •OH and SO4•− with AAP reactions. The whole and branched rate constants were calculated in a suitable temperature range of 198–338 K and 1 atm by using the KiSThelP program. At 298 K and 1 atm, the total rate constants of •OH and SO4•− with AAP were 3.23 × 109 M−1 s−1 and 4.60 × 1010 M−1 s−1, respectively, considering the diffusion-limited effect. The chronic toxicity showed that the main degradation intermediates were harmless to three aquatic organism, namely, fish, daphnia, and green algae. From point of view of the acute toxicity, some degradation intermediates were still at harmful or toxic level. These results provide theoretical guidance on the practical degradation of AAP in the water.

Graphical Abstract

1. Introduction

The problem of water pollution caused by the drug residues have been paid much more attention. Even though the content of these drugs in the water environment is very low, they brings potential dangers to human health and ecological environment due to its strong persistence, bioaccumulation and slow biodegradation [1,2]. Acetaminophen (AAP), as one kind of antipyretic analgesics, enters to water environment by the excretion of humans and animals. The concentration of AAP rose to 6 μg/L in European STP effluents [3]. Its concentration up to 10 μg/L was detected in natural waters in the United States [4], and over 65 μg/L concentration was measured in the Tyne river in the United Kingdom [5]. The removal of micropollution is challenging for water treatment technology. Researches on this area were also relatively limited.
Advanced oxidation processes (AOPs) are highly efficient engineering technologies in the elimination of water micropollutants. The active free radicals (i.e., •OH and SO4•−) generating in AOPs can decompose these contaminants. The ultraviolet/hydrogen peroxide (UV/H2O2), Fenton (H2O2/Fe2+) and Photo-Fenton (UV/H2O2/Fe2+) processes can contribute to •OH, which is shown in the following equation:
Fe2+ + H2O2 → Fe3+ + •OH + HO
•OH could be generated from hydrogen peroxide activated by electrochemical process with Fe2+/Fe3+, which can degrade pentachlorophenol (PCP) [6]. The removal of carbamazepine (CBZ) was attributed to •OH formed by coupling H2O2 with UV and Fe2+/Fe3+ [7]. The removal efficiency of three AOP systems has been compared and found the order of O3/H2O2/Fe2+ > UV/H2O2/Fe2+ > H2O2/Fe2+ [8]. Certainly, other processes such as photocatalysis and photoelectrocatalysis are able to produce •OH [9].
Meanwhile, the ultraviolet/persulfate (UV/S2O82−) can generate SO4•−, which is described by the following equation [10]:
S 2 O 8 2 h v   SO 4 +   SO 4
Surely, •OH can be produced when SO4•− reacts with H2O, which is presented by the following equation [11]:
SO 4 +   H 2 O     SO 4 2 + OH   +   H +
Thus, •OH-mediated and SO4•−-mediated degradation of contaminants were available. The redox potential and rate constants are summarized in Table 1. It was reported that the •OH-initiated degradation rate constant was about 108–1010 M−1 s−1. The SO4•−-initiated rate constant was about 107–1010 M−1 s−1 [12]. The degradation rates of two reactive radicals are nearly equivalent, which is consistent with their high redox potential (2.5–3.1 V for SO4•− versus 1.8–2.7 V for •OH) [13,14,15]. The degradation processes of contaminants triggered by •OH and SO4•− were investigated in recent years. For example, Tong et al. determined the rate constants of syringic acid reactions with •OH and SO4•− in aqueous phase by laser flash photolysis. They found that •OH and SO4•− possessed similar reaction rate at the same pH [16]. Gao et al. measured the rate constants of neutral sulfamethoxazole with •OH and SO4•− were (7.27 ± 0.43) × 109 and (2.98 ± 0.32) × 109 M−1 s−1 in the systems of UV/H2O2 and UV/PS, respectively [17]. Similarly, Wang et al. detected the rate constants for AAP with •OH and SO4•− reactions were (3.26 ± 0.41) × 109 and (1.80 ± 0.17) × 109 M−1 s−1 in the Fe2+/persulfate system, respectively [18]. The second-order rate constants of •OH and SO4•− were conformed as 5.15 × 109 and 7.66 × 109 M−1 s−1, respectively, using the ultraviolet light emitting diode (UV-LED)-based method by Li et al. [19]. However, the study of degradation mechanisms of •OH and SO4•− with the target contaminants still faced with great challenge. At atom level, quantum chemistry calculation was a powerful tool to gain a in-depth understanding for mechanisms and kinetics of •OH and SO4•− reacting with some pollutants [20,21,22,23].
Theoretical studies are essential for discussing the degradation processes of AAP with •OH and SO4•−. Therefore, the reaction mechanisms and kinetics of the AAP with •OH and SO4•− have been studied by using quantum chemistry calculations. Rate constants of every possible pathways for AAP with •OH and SO4•− reactions were calculated. Even more importantly, the ecotoxicity of AAP and its degradation products has been evaluated in order to know their risk.

2. Computational Methods

2.1. Mechanism Calculation

Usually, reaction mechanisms are investigated by Density functional theory (DFT). M06-2X functional can solve noncovalent interactions for some complexes better than other density functional such as B3LYP [24]. The functional ratio of exchange correction of M06-2X is 54% which will obtain more accurate energies [25]. Furthermore, M06-2X method [26] of DFT was selected in the reactions of AAP with •OH and SO4•− without hesitation, because satisfactory results were acquired on the degradation of some micropollutants [27,28,29,30]. For example, the thermodynamic and kinetic data for ibuprofen reactions with hydroxyl and sulfate radicals reported by Yang et al. were calculated with M06-2X method [27]. All electronic structures and energy calculations were performed using Gaussian 09 software [31]. The reactants (R), transition states (TS) and intermediates (IM) were optimized at the M06-2X/6-31+G(d,p) level. IM (all positive frequencies) and TS (only one imaginary frequency) are primarily identified by harmonic vibration frequency analysis. Moreover, the method of intrinsic reaction coordinates (IRC) was applied to determine every right transition state [32]. The water solvent effect was taken into account by a universal solvation model (SMD) [33] when these structures were optimized in the aqueous phase. Based on right structures, the single point energies were calculated accurately at high level of M06-2X/6-311++G(3df, 2p).The Gibbs free energy barrier of activation (ΔG) and free energy of reaction (ΔG) are calculated as follows:
ΔG = GTSGR
ΔG = GIMGR

2.2. Kinetic Calculation

The conventional Transition State Theory (TST) was used to calculate the rate constants implemented by KiSThelP program [34] that has obtained accurate results for contaminants with free radicals reactions [35,36,37,38,39]. The calculation formula is employed in KiSThelP:
k = κ σ k b T h ( R T P 0 ) Δ n e Δ G 0 , k b T
Some parameters need to be explained. κ is tunneling correction factor of Wigner approach [34]. σ, kb and h are the reaction path degeneracy, Boltzmann’s constant and Planck’s constant, respectively. ∆G0,≠ is the standard Gibbs free energy of activation. RT/P0 has the unit of the inverse of a concentration. For bimolecular reactions, ∆n is equal to 1.
The diffusion-limited effect was considered to obtain the apparent rate constants (kapp) of aqueous phase based on Collins-Kimball theory [40].
k a p p = k a q k D k a q + k D
where, kaq is calculated by TST as aqueous rate constant. kD is calculated by the Smoluchowski equation as the diffusion-limited rate constants:
k D = 4 π R A B D A B N A
RAB means the reaction distance, and NA is Avogadro’s number, DAB represents the sum of diffusion coefficient of the reactants A (AAP) and B (•OH or SO4•−). The calculations of DA and DB are realized by using the Stokes–Einstein approach [41]:
D = k b T 6 π α η
where kb, T, η, and α are the Boltzmann constant, temperature, viscosity of the solvent, and radius of the solute, respectively. For water, η = 8.9 × 10−4 Pa s.

2.3. Ecotoxicity Calculation

The aquatic toxicity of AAP and its degradation products was evaluated by using the Structure Activity Relationship (SAR) method with the ECOSAR program [42], which has been successfully used to evaluate the acute and chronic toxicity [43,44,45,46]. Three aquatic organisms of green algae, daphnia and fish were considered to assess the acute and chronic toxicities. Acute toxicity of the target compounds was estimated by median lethal concentration (LC50) and median effect concentration (EC50). LC50 is defined 50% lethal concentration for fish and daphnia in 96 and 48 h, respectively. EC50 is 50% effective concentration for green algae in 96 h. The chronic toxicity was described by the chronic toxicity value (ChV) for green algae, daphnia and fish.

3. Results and Discussion

3.1. Degradation Mechanisms

The degradation mechanisms of AAP initiated by •OH and SO4•− mainly include radical adduct formation (RAF) and formal hydrogen atom transfer (FHAT). Similarities and differences of mechanisms about two reactions were adequately investigated. The Gibbs free energy of reaction (ΔG) and Gibbs free energy barrier of activation (ΔG) of the initial reaction of AAP with •OH and SO4•− were calculated and discussed. The binding distances and angles of AAP, •OH and SO4•− are shown in Figure 1. All structures of transition states are plotted in Figure S1 (Supplementary Materials).

3.1.1. Radical Adduct Formation

RAF pathways of AAP with •OH and SO4•− reactions are displayed in Figure 2. •OH-initiated and SO4•−-initiated RAF channels consist of addition on the benzene ring and the acetamide group. It is uniform for the RAF mechanisms of AAP with •OH and SO4•− reactions. The acetamide group addition has no advantage over that of the benzene ring because the free energy barriers are 15.23 and 30.86 kcal/mol for acetamide group addition of •OH and SO4•−, respectively. However, the free energy barriers were 5.95–9.26 kcal/mol and 2.66–8.74 kcal/mol for •OH and SO4•− addition to six carbon atoms of benzene ring, respectively. Generally, •OH-triggered reactions are higher exothermic than that of SO4•−. Based on the values of ΔG and ΔG, C6 atom addition (path 6 for •OH-triggered reactions versus path 13 for SO4•−-triggered reactions) is the most favorable channels because their barriers are only 5.95 and 2.66 kcal/mol, respectively. Recently, the similar addition results were proved by Li et al. [47]. Figure 3 shows the comparison of potential energies for RAF mechanisms of two radicals reactions. SO4•−-initiated reactions have the lower free energy barriers than that of •OH. In SO4•−-initiated reactions, TS13, C6 addition transition state, has stronger hydrogen bond intermolecular interaction, namely, hydrogen atom of phenolic hydroxyl group of AAP with oxygen atom of SO4•−. IRC intuitively shows hydrogen atom of phenolic hydroxyl group of AAP tends to be close to oxygen atom of SO4•−. The interaction will greatly decrease reaction barrier. However, SO4•−-initiated reactions have less reaction heats compared with •OH-initiated reactions. For example, the energy barrier of path 6 is higher 3.29 kcal/mol than path 13, but path 6 is more exothermic than 6.95 kcal/mol.

3.1.2. Formal Hydrogen Atom Transfer

Ten hydrogen abstraction pathways from benzene ring and methyl group are found and shown in Figure 4. Hydrogen abstractions from C2, C3, and C5 of benzene ring experience TS15, TS16, TS17 with free energy barriers of 19.13, 18.75 and 19.75 kcal/mol for AAP with •OH reactions, respectively. For SO4•−-initiated reactions, the free energy barriers of hydrogen abstractions from C2, C3, C5, and C6 of benzene ring are 30.58, 19.60, 19.30 and 30.43 kcal/mol, respectively. The hydrogen atom can be abstracted from C6 of benzene ring and the methyl group via 14.27 and 14.41 kcal/mol barriers for •OH-initiated path 18 and path 19. Moreover, the corresponding products are exothermic 7.02 and 11.37 kcal/mol, respectively. The results indicate that hydrogen abstractions from C6 of benzene ring and methyl group are two important channels for •OH with AAP reaction. However, methyl group hydrogen abstraction is the most important channel for SO4•− with AAP reaction due to the lowest energy barrier of 10.91 kcal/mol. Figure 5 describes the comparison of free energies for FHAT mechanisms. As shown in the Figure 5, the free energy barriers for SO4•− abstracting hydrogen atom from benzene ring are higher than that of •OH, and the corresponding paths (path 20–path 23) initiated by SO4•− are less exothermic than path 15–path 18 initiated by •OH. However, the free energy barrier of SO4•−-initiated path 24 is lower 3.5 kcal/mol than •OH-initiated path 19, and path 24 is more exothermic 1.76 kcal/mol than path 19.
The comparison of FHAT and RAF mechanisms is shown in Figure 6. It is implied that RAF has an advantage over FHAT for both reactions. The free energy barrier for the most important RAF channel is lower 8.32 and 8.25 kcal/mol than the most favorable FHAT channel for •OH-initiated and SO4•−-initiated reactions, respectively.

3.2. Kinetics

The rate constants involved free radicals reactions are of great value for predicting the degradation rate. However, the measurement of such data is difficult due to these rapid reactions. The theoretical calculations play an important role in attaining rate constants for these radical-participating reactions.
The rate constants for AAP with •OH reactions are given in Table 2. The apparent rate constant of •OH reaction with AAP (kapp) is 3.23 × 109 M–1 s–1 at 298 K. The calculated rate constant is consistent with experimental results of (3.26 ± 0.41) × 109 and 5.15 × 109 M–1 s–1 [18,19]. The C6 site addition (path 6) has the largest rate constant of 3.56 × 109 M–1 s–1 with the 84.8% branching ratio. The other RAF and FHAT pathways contribute the total reactions weakly. The rate constants for AAP with SO4•− reactions are depicted in Table 3. The apparent rate constant of SO4•− reaction with AAP (k’app) is 4.60 × 1010 M–1 s–1 at 298 K, which is higher six times than experimental value of 7.66 × 109 M–1 s–1 [19]. The possible reason is that the lower barrier leads to higher reaction rate, which agrees with discussion of mechanisms. Theoretical model and method will lead to some deviations, but the accuracy of experiment is affected by some factors such as equipment, reagent, and operation. Theoretical calculations can predict and explain some results. Consequently, the benefits of theoretical calculations cannot be underestimated. The C6 site addition (path 13) is dominant channel with the largest rate constant of 8.65 × 1013 M–1 s–1 that possesses the 92.8% branching ratio. The other RAF and FHAT pathways have a little contribution for AAP with SO4•− reaction. As shown in Table 4, C6 of benzene ring and methyl group hydrogen abstractions are dominant channels for •OH with AAP reaction with the branching ratio of 50.42% and 49.58%, respectively. For AAP with SO4•− reaction, methyl group hydrogen abstraction contributes 100% to FHAT channels.
The temperature dependence of rate constants is shown in Figure 7 at the temperatures from 198 to 338 K and 1 atm, and the corresponding data are listed in Tables S1 and S2 (Supplementary Materials). The total rate constants have weakly negative temperature dependence for •OH-initiated reaction. However, SO4•−-initiated reactions have distinctly negative temperature dependence.

3.3. The Aquatic Toxicities of AAP and Its Degradation Intermediates

The acute and chronic toxicities of AAP and the important degradation intermediates are assessed in three different aquatic organisms, which is drawn in Figure 8. Four types are classified and listed in Table S3 (Supplementary Materials). The toxic values of AAP and the important degradation intermediates are shown in Table S4 (Supplementary Materials).

3.3.1. Toxicity of AAP

The acute toxicity value of AAP is calculated as 323 mg/L of LC50 for fish, 63.1 mg/L of LC50 for daphnia and 26.3 mg/L of EC50 for green algae, respectively. These results indicate that AAP is harmful to daphnia and green algae, but not harmful to fish. The calculated ChV of AAP is 26.3 mg/L for fish, 5.13 mg/L for daphnia, and 37.2 mg/L for green algae. AAP is not harmful to fish and green algae at chronic level. However, it is harmful to daphnia chronically.

3.3.2. Toxicities of the Degradation Products

The most important intermediate (IM6) is harmful to three aquatic organisms in acute toxicity, but is harmless to three aquatic organisms in chronic toxicity. For other degradation intermediates, IM1 is acutely toxic for fish and green algae, and harmful to daphnia. The chronic toxicity of IM1 is harmful for three aquatic organisms. IM13 and IM8 are not harmful for three aquatic organisms chronically. Moreover, IM13 and IM8 are not acutely harmful for fish and daphnia, but pose a severe threat for green algae. In brief, the most important degradation intermediate (IM6) from •OH-initiated reaction is still harmful to aquatic organisms. IM13 from SO4•−-initiated reaction is harmless to fish and daphnia, but is very toxic to green algae. Thus, the toxicity of these compounds should be concerned.

4. Conclusions

In this work, the reaction mechanisms and rate of AAP with •OH and SO4•− have been explored theoretically in aqueous phase. The toxicity of AAP and its transformation intermediates to three aquatic organisms have been assessed. The novelty are summarized as below:
(1)
M06-2X/6-311+G (3df, 2p)//M06-2X/6-31+G (d, p) has been used to study the •OH-initiated and SO4•−-initiated transformation mechanism of AAP. •OH and SO4•− with AAP reactions have the same reaction sites, even reaction mechanisms. The results implied that the C6 addition is prominent pathway in RAF mechanisms and hydrogen abstraction of methyl group is dominant pathway for both reactions in FHAT mechanism. RAF takes precedence over FHAT.
(2)
At 298 K, the total apparent rate constant of AAP with SO4•− is larger than that of •OH. The calculated rate constants basically matched with experimental values. Theoretical calculations predicted the kinetic data at 198 K–338 K.
(3)
Toxic assessment shows that some representative degradation intermediates present an acute threat to the target organisms. Thus, subsequent degradation should be implemented until they are degraded into non-toxic substances.
In brief, this work explains the degradation processes of AAP initiated by •OH and SO4•− from microscopic points, and solves the problem of structures of intermediates and products which are associated with reactivity. The calculation of eco-toxicity plays an important role on assessing toxicity of degradation process. Finally, these results can apply to the practical degradation of AAP in AOPs.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/toxics9100234/s1, Figure S1: Optimized geometries involving the transition states of AAP with •OH and SO4•− at the M06-2X/6-31+G(d,p) level, Table S1: Calculated rate constants of AAP with •OH from 198 to 338 K and 1 atm, Table S2: Calculated rate constants of AAP with SO4•− from 198 to 338 K and 1 atm, Table S3: The acute and chronic toxicity class (mg L−1), Table S4: Eco-toxicity values of AAP and its transformation intermediates to aquatic organisms (mg L−1)

Author Contributions

Conceptualization, M.X. and J.Y.; methodology, J.Y.; software, J.S.; validation, S.S., S.Y. and J.S.; formal analysis, M.X.; investigation, M.X.; resources, J.S.; data curation, S.Y.; writing—original draft preparation, M.X.; writing—review and editing, J.Y. and S.S.; visualization, J.Y.; supervision, J.S; project administration, J.S; funding acquisition, J.S. 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, grant number 21507027, and the Natural Science Foundation of Hubei Province, grant number 2020CFB468.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Akhtar, J.; Amin, N.A.S.; Shahzad, K. A review on removal of pharmaceuticals from water by adsorption. Desalin. Water Treat. 2015, 57, 12842–12860. [Google Scholar] [CrossRef]
  2. Papageorgiou, M.; Kosma, C.; Lambropoulou, D. Seasonal occurrence, removal, mass loading and environmental risk assessment of 55 pharmaceuticals and personal care products in a municipal wastewater treatment plant in Central Greece. Sci. Total Environ. 2016, 543, 547–569. [Google Scholar] [CrossRef]
  3. Ternes, T.A. Occurrence of drugs in German sewage treatment plants and rivers. Water Res. 1998, 32, 3245–3260. [Google Scholar] [CrossRef]
  4. Kolpin, D.W.; Furlong, E.T.; Meyer, M.T.; Thurman, E.M.; Zaugg, S.D.; Barber, L.B.; Buxton, H.T. Pharmaceuticals, hormones, and other organic wastewater contaminants in U.S. streams, 1999–2000: A national reconnaissance. Environ. Sci. Technol. 2002, 36, 1202–1211. [Google Scholar] [CrossRef] [Green Version]
  5. Roberts, P.H.; Thomas, K.V. The occurrence of selected pharmaceuticals in wastewater effluent and surface waters of the lower Tyne catchment. Sci. Total Environ. 2006, 356, 143–153. [Google Scholar] [CrossRef] [PubMed]
  6. Govindan, K.; Raja, M.; Noel, M.; James, E.J. Degradation of pentachlorophenol by hydroxyl radicals and sulfate radicals using electrochemical activation of peroxomonosulfate, peroxodisulfate and hydrogen peroxide. J. Hazard. Mater. 2014, 272, 42–51. [Google Scholar] [CrossRef] [PubMed]
  7. Ali, F.; Khan, J.A.; Shah, N.S.; Sayed, M.; Khan, H.M. Carbamazepine degradation by UV and UV-assisted AOPs: Kinetics, mechanism and toxicity investigations. Process. Saf. Environ. 2018, 117, 307–314. [Google Scholar] [CrossRef]
  8. Khodadadi, T.; Solgi, E.; Mortazavi, S.; Nourmoradi, H. Comparison of advanced oxidation methods of Fenton, UV/Fenton, and O3/Fenton in treatment of municipal wastewater. Desalin. Water Treat. 2020, 206, 108–115. [Google Scholar] [CrossRef]
  9. Collivignarelli, M.C.; Miino, M.C.; Arab, H.; Bestetti, M.; Franz, S. Efficiency and energy demand in polishing treatment of wastewater treatment plants effluents: Photoelectrocatalysis vs. Photocatalysis and photolysis. Water 2021, 13, 821. [Google Scholar] [CrossRef]
  10. Moussavi, G.; Pourakbar, M.; Aghayani, E.; Mahdavianpour, M. Investigating the aerated VUV/PS process simultaneously generating hydroxyl and sulfate radicals for the oxidation of cyanide in aqueous solution and industrial wastewater. Chem. Eng. J. 2018, 350, 673–680. [Google Scholar] [CrossRef]
  11. Zhang, Y.C.; Zhang, Q.; Hong, J.M. Sulfate radical degradation of acetaminophen by novel iron–copper bimetallic oxidation catalyzed by persulfate: Mechanism and degradation pathways. Appl. Surf. Sci. 2017, 422, 443–451. [Google Scholar] [CrossRef]
  12. Li, M.X.; Sun, J.F.; Han, D.D.; Wei, B.; Mei, Q.; An, Z.X.; Wang, X.Y.; Cao, H.J.; Xie, J.; He, M.X. Theoretical investigation on the contribution of HO•, SO4•− and CO3•− radicals to the degradation of phenacetin in water: Mechanisms, kinetics, and toxicity evaluation. Ecotoxicol. Environ. Saf. 2020, 204, 110977. [Google Scholar] [CrossRef]
  13. Xiao, R.Y.; He, L.; Luo, Z.H.; Spinney, R.; Wei, Z.S.; Dionysiou, D.D.; Zhao, F.P. An experimental and theoretical study on the degradation of clonidine by hydroxyl and sulfate radicals. Sci. Total Environ. 2020, 710, 136333. [Google Scholar] [CrossRef]
  14. Devi, P.; Das, U.; Dalai, A.K. In-situ chemical oxidation: Principle and applications of peroxide and persulfate treatments in wastewater systems. Sci. Total Environ. 2016, 571, 643–657. [Google Scholar] [CrossRef]
  15. Ghanbari, F.; Moradi, M. Application of peroxymonosulfate and its activation methods for degradation of environmental organic pollutants: Review. Chem. Eng. J. 2017, 310, 41–62. [Google Scholar] [CrossRef]
  16. Tong, X.; Wang, S.N.; Wang, L.M. Kinetics and mechanism of syringic acid degradation initiated by hydroxyl radical and sulphate radical in the aqueous phase. Chemosphere 2020, 256, 126997. [Google Scholar] [CrossRef]
  17. Gao, L.W.; Mao, Q.M.; Luo, S.; Cao, L.Y.; Xie, X.D.; Yang, Y.; Deng, Y.F.; Wei, Z.S. Experimental and theoretical insights into kinetics and mechanisms of hydroxyl and sulfate radicals-mediated degradation of sulfamethoxazole: Similarities and differences. Environ. Pollut. 2020, 259, 113795. [Google Scholar] [CrossRef]
  18. Wang, S.L.; Wu, J.F.; Lu, X.Q.; Xu, W.X.; Gong, Q.; Ding, J.Q.; Dan, B.S.; Xie, P.C. Removal of acetaminophen in the Fe2+/persulfate system: Kinetic model and degradation pathways. Chem. Eng. J. 2019, 358, 1091–1100. [Google Scholar] [CrossRef]
  19. Li, B.Q.; Ma, X.Y.; Deng, J.; Li, Q.S.; Chen, W.Z.; Li, G.X.; Chen, G.Y.; Wang, J.P. Comparison of acetaminophen degradation in UV-LED-based advance oxidation processes: Reaction kinetics, radicals contribution, degradation pathways and acute toxicity assessment. Sci. Total Environ. 2020, 723, 137993. [Google Scholar] [CrossRef] [PubMed]
  20. Li, H.X.; Miao, X.R.; Zhang, J.; Du, J.; Xu, S.D.; Tang, J.H.; Zhang, Y.Y. DFT studies on the reaction mechanism and kinetics of dibutyl phthalate initiated by hydroxyl and sulfate radicals: Prediction of the most reactive sites. Chem. Eng. J. 2020, 381, 122680. [Google Scholar] [CrossRef]
  21. Liu, W.; Lv, G.C.; Sun, X.M.; He, L.; Zhang, C.X.; Li, Z.Q. Theoretical study on the reaction of anthracene with sulfate radical and hydroxyl radical in aqueous solution. Ecotoxicol. Environ. Saf. 2019, 183, 109551. [Google Scholar] [CrossRef]
  22. Mei, Q.; Sun, J.F.; Han, D.D.; Wei, B.; An, Z.X.; Wang, X.Y.; Xie, J.; Zhan, J.H.; He, M.X. Sulfate and hydroxyl radicals-initiated degradation reaction on phenolic contaminants in the aqueous phase: Mechanisms, kinetics and toxicity assessment. Chem. Eng. J. 2019, 373, 668–676. [Google Scholar] [CrossRef]
  23. Yang, J.X.; Lv, G.C.; Wang, Z.H.; Sun, X.M.; Gao, J. Mechanisms, kinetics and eco-toxicity assessment of singlet oxygen, sulfate and hydroxyl radicals-initiated degradation of fenpiclonil in aquatic environments. J. Hazard. Mater. 2021, 409, 124505. [Google Scholar] [CrossRef] [PubMed]
  24. Hohenstein, E.G.; Chill, S.T.; Sherrill, C.D. Assessment of the performance of the M05-2X and M06-2X exchange-correlation functionals for noncovalent interactions in biomolecules. J. Chem. Theory Comput. 2008, 4, 1996–2000. [Google Scholar] [CrossRef] [PubMed]
  25. Mardirossian, N.; Head-Gordon, M. Thirty years of density functional theory in computational chemistry: An overview and extensive assessment of 200 density functionals. Mol. Phys. 2017, 115, 2315–2372. [Google Scholar] [CrossRef]
  26. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef] [Green Version]
  27. Yang, Z.H.; Su, R.H.; Luo, S.; Spinney, R.; Cai, M.Q.; Xiao, R.Y.; Wei, Z.S. Comparison of the reactivity of ibuprofen with sulfate and hydroxyl radicals: An experimental and theoretical study. Sci. Total. Environ. 2017, 590, 751–760. [Google Scholar] [CrossRef]
  28. He, L.; Sun, X.M.; Zhu, F.P.; Ren, S.J.; Wang, S.G. OH-initiated transformation and hydrolysis of aspirin in AOPs system: DFT and experimental studies. Sci. Total Environ. 2017, 592, 33–40. [Google Scholar] [CrossRef]
  29. Yao, J.F.; Tang, Y.Z.; Zhang, Y.J.; Ruan, M.; Chen, F.; Wu, W.Z.; Sun, J.Y. Quantum chemical study of the mechanisms, kinetics, and ecotoxicity assessment of OH radical-initiated reactions of 2, 2′, 4, 4′, 5, 5′-hexabrominated diphenyl ether (BDE-153) in atmosphere and wastewater. Chem. Eng. J. 2021, 422, 129916. [Google Scholar] [CrossRef]
  30. Zou, M.T.; Qi, Y.M.; Qu, R.J.; Al-Basher, G.; Pan, X.X.; Wang, Z.Y.; Huo, Z.L.; Zhu, F. Effective degradation of 2,4-dihydroxybenzophenone by zero-valent iron powder (Fe0)-activated persulfate in aqueous solution: Kinetic study, product identification and theoretical calculations. Sci. Total Environ. 2021, 771, 144743. [Google Scholar] [CrossRef]
  31. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, V.; Barone, G.A.; Petersson, H.; Nakatsuji, X.; et al. Gaussian 16 (Revision A.03); Gaussian, Inc.: Pittsburgh, PA, USA, 2016. [Google Scholar]
  32. Maeda, S.; Harabuchi, Y.; Ono, Y.; Taketsugu, T.; Morokuma, K. Intrinsic reaction coordinate: Calculation, bifurcation, and automated search. Int. J. Quant. Chem. 2015, 115, 258–269. [Google Scholar] [CrossRef]
  33. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal solvation model based on solute electron density and on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef] [PubMed]
  34. Canneaux, S.; Bohr, F.; Henon, E. KiSThelP: A program to predict thermodynamic properties and rate constants from quantum chemistry results. J. Comput. Chem. 2014, 35, 82–93. [Google Scholar] [CrossRef] [PubMed]
  35. Altarawneh, M.; Dlugogorski, B.Z. Formation and chlorination of carbazole, phenoxazine, and phenazine. Environ. Sci. Technol. 2015, 49, 2215–2221. [Google Scholar] [CrossRef] [Green Version]
  36. Wang, X.Y.; Sun, J.F.; Han, D.D.; Bao, L.; Mei, Q.; Wei, B.; An, Z.X.; He, M.X.; Yuan, S.L.; Xie, J.; et al. Gaseous and heterogeneous reactions of low-molecular-weight (LMW) unsaturated ketones with O3: Mechanisms, kinetics, and effects of mineral dust in tropospheric chemical processes. Chem. Eng. J. 2020, 395, 125083. [Google Scholar] [CrossRef]
  37. Yao, J.F.; Sun, Y.N.; Tang, Y.Z.; Zhang, Y.J.; Wu, W.Z.; Sun, J.Y. Atmospheric oxidation of 4-(2-methoxyethyl) phenol initiated by OH radical in the presence of O2 and NOx: A mechanistic and kinetic study. Int. J. Quantum. Chem. 2021, 121, e26650. [Google Scholar] [CrossRef]
  38. An, Z.X.; Sun, J.F.; Han, D.D.; Mei, Q.; Wei, B.; Wang, X.Y.; He, M.X. Theoretical study on the mechanisms, kinetics and ecotoxicity assessment of OH-initiated reactions of guaiacol in atmosphere and wastewater. Sci. Total Environ. 2019, 685, 729–740. [Google Scholar] [CrossRef]
  39. Ji, Y.M.; Zhao, J.; Terazono, H.; Misawa, K.; Levitt, N.P.; Li, Y.X.; Lin, Y.; Peng, J.F.; Wang, Y.; Duan, L.; et al. Reassessing the atmospheric oxidation mechanism of toluene. Proc. Natl. Acad. Sci. USA 2017, 114, 8169–8174. [Google Scholar] [CrossRef] [Green Version]
  40. Collins, F.C.; Kimball, G.E. Diffusion-controlled reaction rates. J. Colloid Sci. 1949, 4, 425–437. [Google Scholar] [CrossRef]
  41. Steubing, W. Ueber die von der molekularkinetischen Theorie der Wärme geforderte Bewegung von in Flüssigkeiten suspendierten Teilchen. Z. Chem. Ind. Kolloide 1908, 3, 230. [Google Scholar] [CrossRef]
  42. ECOSAR v2.0. Available online: https://www.epa.gov/oppt/newchems/tools/21ecosar.htm (accessed on 1 January 2017).
  43. Bai, F.Y.; Ni, S.; Tang, Y.Z.; Pan, X.M.; Zhao, Z. Ciprofloxacin transformation in aqueous environments: Mechanism, kinetics, and toxicity assessment during OH-mediated oxidation. Sci. Total. Environ. 2020, 699, 134190. [Google Scholar] [CrossRef]
  44. Shada, A.; Chen, J.; Qu, R.J.; Dar, A.A.; Bin-Jumah, M.; Allam, A.A.; Wang, Z.Y. Degradation of sulfadimethoxine in phosphate buffer solution by UV alone, UV/PMS and UV/H2O2: Kinetics, degradation products, and reaction pathways. Chem. Eng. J. 2020, 398, 125357. [Google Scholar] [CrossRef]
  45. Chen, J.; Xu, X.X.; Zeng, X.L.; Feng, M.B.; Qu, R.J.; Wang, Z.Y.; Nesnas, N.; Sharma, V.K. Ferrate(VI) oxidation of polychlorinated diphenyl sulfides: Kinetics, degradation, and oxidized products. Water Res. 2018, 143, 1–9. [Google Scholar] [CrossRef] [PubMed]
  46. Shiroudi, A.; Deleuze, M.S.; Canneaux, S. Theoretical study of the oxidation mechanisms of naphthalene initiated by hydroxyl radicals: The O2 addition reaction pathways. Phys. Chem. Chem. Phys. 2015, 17, 13719–13732. [Google Scholar] [CrossRef] [PubMed]
  47. Li, M.X.; Sun, J.F.; Mei, Q.; Wei, B.; An, Z.X.; Cao, H.J.; Zhang, C.; Xie, J.; Zhan, J.H.; Wang, W.X.; et al. Acetaminophen degradation by hydroxyl and organic radicals in the peracetic acid-based advanced oxidation processes: Theoretical calculation and toxicity assessment. J. Hazard. Mater. 2021, 416, 126250. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The structures of AAP, •OH and SO4•− calculated at the M06-2X/6-31 + G(d,p) level. The bond distance (Å) and angles (°) are listed. Here, Toxics 09 00234 i001.
Figure 1. The structures of AAP, •OH and SO4•− calculated at the M06-2X/6-31 + G(d,p) level. The bond distance (Å) and angles (°) are listed. Here, Toxics 09 00234 i001.
Toxics 09 00234 g001
Figure 2. The radical adduct formation channels of AAP with •OH and SO4•− reactions with the Gibbs free energy of reaction (ΔG) and Gibbs free energy barrier of activation (ΔG) (unit: kcal/mol).
Figure 2. The radical adduct formation channels of AAP with •OH and SO4•− reactions with the Gibbs free energy of reaction (ΔG) and Gibbs free energy barrier of activation (ΔG) (unit: kcal/mol).
Toxics 09 00234 g002
Figure 3. The free energy diagram of RAF pathways initiated by two radicals.
Figure 3. The free energy diagram of RAF pathways initiated by two radicals.
Toxics 09 00234 g003
Figure 4. The formal hydrogen atom transfer channels of AAP with •OH and SO4•− reactions with the Gibbs free energy of reaction (ΔG) and Gibbs free energy barrier of activation (ΔG) (unit: kcal/mol).
Figure 4. The formal hydrogen atom transfer channels of AAP with •OH and SO4•− reactions with the Gibbs free energy of reaction (ΔG) and Gibbs free energy barrier of activation (ΔG) (unit: kcal/mol).
Toxics 09 00234 g004
Figure 5. The free energy diagram of FHAT pathways initiated by two radicals in the aqueous phase.
Figure 5. The free energy diagram of FHAT pathways initiated by two radicals in the aqueous phase.
Toxics 09 00234 g005
Figure 6. Free energy profiles for FHAT and RAF mechanisms in aqueous phase. (a) APP with •OH reactions; (b) APP with SO4•− reactions.
Figure 6. Free energy profiles for FHAT and RAF mechanisms in aqueous phase. (a) APP with •OH reactions; (b) APP with SO4•− reactions.
Toxics 09 00234 g006
Figure 7. Temperature dependence of the calculated rate constants.
Figure 7. Temperature dependence of the calculated rate constants.
Toxics 09 00234 g007
Figure 8. Acute and chronic toxicity (mg/L) of AAP and its transformation intermediates to aquatic organisms (fish, daphnia, and green algae).
Figure 8. Acute and chronic toxicity (mg/L) of AAP and its transformation intermediates to aquatic organisms (fish, daphnia, and green algae).
Toxics 09 00234 g008
Table 1. Redox potential and kinetic data for the reactions of •OH and SO4•−.
Table 1. Redox potential and kinetic data for the reactions of •OH and SO4•−.
RadicalsRedox Potential a (V)The Range of Rate Constants b
(M−1 s−1)
The Second-Order Rate Constants of Neutral Sulfamethoxazole c (M−1 s−1)The Second-Order Rate Constants of Acetaminophen (M−1 s−1)
Fe2+/PS dUV-LED/H2O2 e
•OH1.8–2.7108–1010(7.27 ± 0.43) × 109(3.26 ± 0.41) × 1095.15 × 109
SO4•−2.5–3.1107–1010(2.98 ± 0.32) × 109(1.80 ± 0.17) × 1097.66 × 109
a (Xiao, et al., 2020; Devi, et al., 2016; Ghanbari, et al., 2017); b (Li, et al. 2020, [12]); c (Gao, et al., 2020); d (Wang, et al., 2019); e (Li, et al., 2020, [19]).
Table 2. The calculated rate constants(kaq), steady-state rate constant (kD), apparent rate constant (kapp) and branching ratio (Raq) for the AAP with •OH reaction in the aqueous phase at 298 K.
Table 2. The calculated rate constants(kaq), steady-state rate constant (kD), apparent rate constant (kapp) and branching ratio (Raq) for the AAP with •OH reaction in the aqueous phase at 298 K.
Pathskaq (M−1 s−1)Raq (%)kD (M−1 s−1)kapp (M−1 s−1)
APP + •OH → IM1 (k1)8.04 × 1071.99.80 × 1097.97 × 107
APP + •OH → IM2 (k2)1.87 × 1084.59.80 × 1091.83 × 108
APP + •OH → IM3 (k3)1.51 × 1070.49.80 × 1091.51× 107
APP + •OH → IM4 (k4)3.22 × 1087.69.80 × 1093.12 × 108
APP + •OH → IM5 (k5)3.33 × 1070.89.80 × 1093.32 × 107
APP + •OH → IM6 (k6)3.56 × 10984.89.80 × 1092.61 × 109
APP + •OH → IM7 (k7)6.75 × 10209.80 × 1096.75 × 102
APP + •OH → IM15 (k15)4.7509.80 × 1094.75
APP + •OH → IM16 (k16)9.7509.80 × 1099.75
APP + •OH → IM17 (k17)1.9109.80 × 1091.91
APP + •OH → IM18 (k18)1.15 × 10409.80 × 1091.15 × 104
APP + •OH → IM19 (k19)1.13 × 10409.80 × 1091.13 × 104
APP + •OH → Product (ktotal)4.20 × 109100 3.23 × 109
Table 3. The calculated rate constants(k’aq), steady-state rate constant (k’D), apparent rate constant (k’app) and branching ratio (R’aq) for the AAP with SO4•− reaction in the aqueous phase at 298 K.
Table 3. The calculated rate constants(k’aq), steady-state rate constant (k’D), apparent rate constant (k’app) and branching ratio (R’aq) for the AAP with SO4•− reaction in the aqueous phase at 298 K.
Pathsk’aq (M−1 s−1)R’aq (%)k’D (M−1 s−1)k’app (M−1 s−1)
APP + SO4•−→IM8 (k’8)6.00 × 10126.48.05 × 1098.04 × 109
APP + SO4•− → IM9 (k’9)1.61 × 10110.28.05 × 1097.67 × 109
APP + SO4•− → IM10 (k’10)2.60 × 10110.38.05 × 1097.81 × 109
APP + SO4•− → IM11 (k’11)3.28 × 10100.018.05 × 1096.46 × 109
APP + SO4•− → IM12 (k’12)2.52 × 10110.38.05 × 1097.80 × 109
APP + SO4•− → IM13 (k’13)8.65 × 101392.88.05 × 1098.05 × 109
APP + SO4•− → IM14 (k’14)1.77 × 10−608.05 × 1091.77 × 10−6
APP + SO4•− → IM20 (k’20)14.308.05 × 10914.3
APP + SO4•− → IM21 (k’21)1.11 × 10208.05 × 1091.11 × 102
APP + SO4•− → IM22 (k’22)1.88 × 10208.05 × 1091.88 × 102
APP + SO4•− → IM23 (k’23)5.5508.05 × 1095.55
APP + SO4•− → IM24 (k’24)1.33 × 10808.05 × 1091.33 × 108
APP + SO4•− → Product (k’total)9.32 × 1013100 4.60 × 1010
Table 4. The calculated rate constants (kaq, kaq) and branching ratio (Raq, Raq) for the formal hydrogen atom transfer channels in the aqueous phase at 298 K.
Table 4. The calculated rate constants (kaq, kaq) and branching ratio (Raq, Raq) for the formal hydrogen atom transfer channels in the aqueous phase at 298 K.
Pathskaq (M−1 s−1)Raq (%)Pathsk’aq (M−1 s−1)R’aq (%)
APP + •OH (FHAT)2.28 × 104100APP + SO4•− (FHAT)1.33 × 108100
APP + •OH → IM15 (k15)4.750APP + SO4•− → IM20 (k’20)14.30
APP + •OH → IM16 (k16)9.750APP + SO4•− → IM21 (k’21)1.11 × 1020
APP + •OH → IM17 (k17)1.910APP + SO4•− → IM22 (k’22)1.88 × 1020
APP + •OH → IM18 (k18)1.15 × 10450.42APP + SO4•− → IM23 (k’23)5.550
APP + •OH → IM19 (k19)1.13 × 10449.58APP + SO4•− → IM24 (k’24)1.33 × 108100
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xu, M.; Yao, J.; Sun, S.; Yan, S.; Sun, J. Theoretical Calculation on the Reaction Mechanisms, Kinetics and Toxicity of Acetaminophen Degradation Initiated by Hydroxyl and Sulfate Radicals in the Aqueous Phase. Toxics 2021, 9, 234. https://doi.org/10.3390/toxics9100234

AMA Style

Xu M, Yao J, Sun S, Yan S, Sun J. Theoretical Calculation on the Reaction Mechanisms, Kinetics and Toxicity of Acetaminophen Degradation Initiated by Hydroxyl and Sulfate Radicals in the Aqueous Phase. Toxics. 2021; 9(10):234. https://doi.org/10.3390/toxics9100234

Chicago/Turabian Style

Xu, Mengmeng, Junfang Yao, Simei Sun, Suding Yan, and Jingyu Sun. 2021. "Theoretical Calculation on the Reaction Mechanisms, Kinetics and Toxicity of Acetaminophen Degradation Initiated by Hydroxyl and Sulfate Radicals in the Aqueous Phase" Toxics 9, no. 10: 234. https://doi.org/10.3390/toxics9100234

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