Next Article in Journal
Phytochemicals and Biological Activities of Garcinia atroviridis: A Critical Review
Previous Article in Journal
Toxicological Findings of Self-Poisoning Suicidal Deaths: A Systematic Review by Countries
Previous Article in Special Issue
Homogeneous and Heterogeneous Photocatalysis for the Treatment of Pharmaceutical Industry Wastewaters: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of pH on the Kinetics and Products of Photocatalytic Degradation of Sulfonamides in Aqueous Solutions

Department of General and Analytical Chemistry, Faculty of Pharmaceutical Sciences in Sosnowiec, Medical University of Silesia in Katowice, Jagiellońska 4, 41-200 Sosnowiec, Poland
*
Author to whom correspondence should be addressed.
Toxics 2022, 10(11), 655; https://doi.org/10.3390/toxics10110655
Submission received: 28 September 2022 / Revised: 19 October 2022 / Accepted: 26 October 2022 / Published: 29 October 2022
(This article belongs to the Special Issue Photocatalytic Degradation of Pharmaceuticals in Water)

Abstract

:
The aims of the study were to determine the kinetics of the photocatalytic degradation of six sulfonamides in the presence of TiO2-P25 in acidic, neutral, and alkaline solutions and to identify the structures of the stable products. It was stated that the pH of the solution significantly affected the photocatalytic degradation rate of sulfonamides in acidic and alkaline environments, and the effect likely depended on the susceptibility of sulfonamides to attack by hydroxyl radicals. In the post-reaction mixture, we identified the compounds resulting from the substitution of the aromatic rings with a hydroxyl group; the amide hydrolysis products; the hydroxylamine-, azo, and nitro derivatives; and the compounds formed via the elimination of the sulfone group. Moreover, previously unknown azo compounds were detected. Some degradation products of sulfonamides may exhibit marked bacteriostatic activity and high phytotoxicity. The azo and nitro compounds formed in an acidic environment may be potentially more toxic to aquatic ecosystems than the initial compounds.

1. Introduction

Sulfonamides (SNs), sulfanilamide derivatives (SADs), are antimicrobial drugs widely used particularly in veterinary medicine. More than 70% of global antimicrobial use has been estimated to be in livestock. In 2020, the antibiotics sales for food-producing animals in 31 European countries and the United States reached more than 5500 t and 10,500 t, of which SNs were 9.9% and 7.2%, respectively [1,2]. After use, both the unmetabolized SNs as well as their metabolites are released directly into the environment. These drugs are quite persistent and nonbiodegradable. Therefore, a significant proportion of them remains biologically active in the environment over a long time [3]. For example, almost 50 mg of SNs per kg of dry weight was detected in the manure from livestock farms [4]. In recent years, only a slight decrease in SN consumption has been observed [1,2]. SAD derivatives are also introduced into the environment as herbicides. Although in 2011 the European Commission decided not to re-register Asulam (methyl (4-aminobenzene-1-sulfonyl) carbamate), it is still manufactured and used in many countries [5,6,7,8,9].
SNs exhibit bacteriostatic effects, but their toxic concentrations to microorganisms are relatively high. According to Ferrari et al. [10], the determined EC50 values of sulfamethoxazole (SMX) against the Vibrio fischeri strain ranged from 43 to 53 mg/L. In turn, the IC50 value of SMX against Arthrobacter globiformis was >127 mg/L [11]. The sensitivity of invertebrates and vertebrates to SMX is also not high [10,12,13,14]. On the other hand, the EC(IC)50 values of SMX against algae (Scenedesmus vacuolatus) were about 1.5 mg/L, and for higher plants (Daucus carota), even <0.05 mg/L [15,16].
These data indicate that SNs are highly phytotoxic. Thiele-Bruhn and Beck reported that soil biocenoses were even more sensitive to this antibiotics group [17]. Another type of risk is related to the long-term exposure of environmental microorganisms to sub-toxic concentrations of SNs. This may result in the emergence of antibiotic-resistant genes (ARGs) and their potential transfer to pathogens [18]. For this reason, the treatment of wastewater containing SNs using biological methods is also risky [19].
An attempt to solve the problem described above is to use a photocatalytic process to remove SNs from wastewater [20,21,22,23]. Hydroxyl (HO) and superoxide (O2−•) radicals are generated during the irradiation of aqueous solutions containing a photocatalyst. The presence of free radicals can be confirmed by analyzing the products of their reaction, e.g., HO with benzene, or using the ESR method [24,25,26,27]. Currently, experiments on extending the useful irradiation spectrum in the heterogenous photocatalysis are carried out [28,29]. They even initiate the decomposition of non-biodegradable organic compounds [30,31,32,33,34,35,36,37,38]. In addition to the effective degradation of antibiotics, the photocatalytic process also affects the inactivation of ARGs, the disinfection of the processed material, and sometimes the complete mineralization of pollutants [34,39]. It was shown that even a short-term irradiation of the solutions in the presence of photocatalyst(s) causes a significant decrease in the concentration of the degraded antibiotic, although its transformation products remain in the post-reaction mixture. Simultaneously, an inhibition of antimicrobial activity in antibiotic solutions was observed [40]. However, the low susceptibility of organisms used in popular toxicity tests to SNs can be a serious problem in the assessment of ecotoxicological effects. As a consequence, in order to observe the toxic effects of the initial solutions of SNs, their concentrations used in the tests must be much higher than those observed in the environment [4]. In addition, the concentrations of the decomposition products of SNs must be high enough to be able to detect their potential toxicity.
Wastewater containing antibiotic residues, including SNs, may be slightly alkaline (livestock wastewater) but may also be close to neutral (municipal wastewater) or acidic (industrial wastewater). This fact should be taken into account when planning the use of photocatalysis for wastewater treatment. The pH has been shown to affect the dynamics of photocatalytic degradation of SNs [35,41,42]. Therefore, pH values likely play a significant role in the degradation pathways of these antibiotics. The effect of pH on free radicals generation during the photocatalytic process in the presence of TiO2 was thoroughly described [38,43,44].
The aims of the study were to determine the kinetics of the photocatalytic degradation of six SNs (Table 1) in the presence of TiO2-P25 in acidic, neutral, and alkaline aqueous solutions and to identify the structures of the stable products. Based on these experiments, the effect of pH on the mechanism of SN degradation and the potential toxicity of the products were assessed.

2. Materials and Methods

2.1. Reagents

The characteristics of the SNs used in the experiments are presented in Table 1.
In addition, TiO2-P25 (Evonik Aeroxide®), HCl and NaOH solution (p.a.; 1 mol/L; POCH), water (LC-MS CHROMASOLV®; Fluka-Analytical, Buchs, Switzerland), acetonitrile (LC-MS CHROMASOLV®; Fluka-Analytical, Buchs, Switzerland), and formic acid (p.a. SIGMA-ALDRICH, St. Louis, MO, USA) were used in the experiments.

2.2. Photocatalytic Process

The stock solutions (0.1 mmol/L) of six SNs (Table 1) were prepared in redistilled water. A 150 mL solution of each SN was placed in a glass beaker with a glass barbotage tube, and 75 mg of TiO2-P25 was added. In the dark, the pH of each sample was adjusted to the expected value with HCl or NaOH solutions. The mixtures in the beakers were irradiated from above with fluorescent lamps (TL-K 40W Actinic BL; Philips). The intensity of irradiation absorbed by the mixtures was 0.37 W/L. Throughout the experiment, the samples were stirred with compressed air.

2.3. Samples Analysis

Aliquots for analysis were removed from the mixtures prior to irradiation and after a specific time. They were filtered (CHROMAFIL RC-20/25; Macherey Nagel) directly into glass vials and immediately analyzed using an Acquity UPLC/DAD system coupled with Xevo G2-XS-QTOF (Waters). The degradation products of SNs were separated using an Acquity UPLC BEH C18 column, 100 × 2.1 mm (Waters), and the mobile gradient phase consisted of a mixture of acetonitrile and water at 35 °C. The DAD detector recorded peaks at 272 nm, while the QTOF detector operated sequentially in ES+ MS and ES+ MS/MS modes. The details of the analysis procedure are presented in Table 2.
The injection volumes of the sample were 1 µL and 5 µL during the quantitative and qualitative analyses, respectively.

2.4. Analysis of the Results

The kinetics of SN degradation were assessed based on the peaks recorded using the DAD detector. For each experiment, a linear regression model of the relationship ln C0/C = f(t) was determined, where the C0/C is the peak area quotient of SNs in the initial solution and in the solution after irradiation time t. This procedure is correct and common for pseudo-first-order reactions.
Molecular formulas of the degradation products of the SNs were determined based on the monoisotopic masses of molecular ions (M+H+) obtained using the MS/TOF technique with ES+ ionization. Structural formulas of the SN degradation products were proposed based on their molecular formulas, and fragmentation spectra were determined using the MS/MS/QTOF technique with ES+ ionization, at collision energies from 10 to 25 V. Aliphatic degradation products of the SNs were not identified.
One of the products of SN transformation is SAD, which is hydrolyzed to sulfanilic acid during ES+ ionization. Therefore, this compound was additionally identified based on the retention time (tR).

2.5. Toxicity Prediction

The potential ecotoxicity of the identified degradation products was estimated only for SMX using ECOSAR (Ecological Structure Activity Relationships) Application 2.2. The ecotoxicity prediction also extended to the aliphatic degradation products of SMX, which were not identified during the chromatographic analysis but are described in the literature.

3. Results

3.1. Kinetics of SN Degradation

In all the performed experiments, the determined relationship lnC0/C = f(t) was the linear function with a coefficient of determination (R2) > 0.9. This indicated that the photocatalytic degradation of SNs could be described using pseudo-first order kinetics. The experimentally found values of the reaction rate constant (k) are shown in Figure 1.
Under the constant experimental conditions, there was no unambiguous and constant relationship between the pH and the degradation rate of SNs. In a neutral environment, the degradation rates of all of the studied SNs were similar, with the exception of SCP. In contrast, in acidic and alkaline environments, the effect varied and depended on the type of SN. Similar observations were already described in previously published articles [41,42].
The photocatalytic degradation of SNs performed in the aqueous environment and in the presence of TiO2 can be initiated by photo holes (h+), HO•0, and depending on the pH, O2−• or hydroperoxide radicals (HO2) [33,34,35,36,37,38,41]. The participation of h+ in a direct reaction with SNs requires the prior adsorption of the reactants onto the photocatalyst surface. The lifetime of HO radicals generated on the TiO2 surface is extremely short (~10−9 s). This indicates that in the case of these radicals, a prior adsorption of the reactants is necessary [41]. Therefore, the photocatalytic reaction rate in the presence of TiO2 suspension should mainly depend on the degree of adsorption of the reactants.
SNs have amphoteric properties (Table 1). However, at a pH of 10–11, the TiO2 surface is negatively charged, and the SNs exist as organic anions. Under these conditions, an electrostatic repulsion between catalyst particles and reagent molecules occurs, and the photocatalytic reaction should be inhibited. A similar phenomenon should be observed in solutions at pH 2–3, where the TiO2 surface is positively charged and SN molecules exist as cations or neutral molecules; however, we observed no such effect.
In our opinion, the degradation rate of SNs in acidic and alkaline environments is primarily determined by the susceptibility of SN molecules to electrophilic attack by HO radicals. The stable products of the direct reaction of SNs with HO identified in acidic solutions (Section 3.2) confirmed our assumptions. Moreover, the lower degradation rate of SNs (except for SCP) in the acidic environment may be associated with the protonation of lone electron pairs in amide and ammonium nitrogen. On the other hand, in an acidic environment, SNs are more susceptible to nucleophilic attack by HO2 radicals. They are much less reactive than HO and have a relatively long half-life but may be responsible for the formation of other products.
Under alkaline conditions, SNs are almost completely dissociated (>99%), with the exception of SAD [46]. A positive linear correlation between the pKa2 values and the degradation rate constant (k) in the presence of TiO2 P25 was determined for five of the six SNs studied (Figure 2). The SMX result was omitted from the regression equation.
The increase in the pKa2 value was a consequence of a lesser electron affinity of the amide group substituent of the studied SNs. Compounds having a lower electron affinity could show higher electron-donor properties. Therefore, this may be the factor that determines the reaction rate of SNs with HO radicals in the alkaline environment and affects the photocatalytic reaction rates. However, in this case, the photodegradation rate with SMX was higher than expected (Figure 2). The probable cause of this phenomenon is the steric effect. SNs differ from each other not only in the heterocyclic ring but also in the arrangement of the spatial structure [42]. It can affect the degradation rate of these compounds in heterogeneous catalysis. Moreover, SMX differs from other SNs in the ability to form hydrogen bonds in the aquatic environment [47].

3.2. Degradation Products

Organic cyclic and aliphatic products, and finally, inorganic compounds were present in the post-reaction solution after the photocatalytic degradation of SNs in the presence of TiO2 P25. The simplest compound of the series of the studied SNs was SAD. Tačić et al. [48] proposed its degradation pathway, and aniline, benzoquinone, sulfanilic acid, and aliphatic compounds such as maleic acid, fumaric acid, oxalic acid, acetic acid, formic acid, and inorganic compounds were identified among the products. In turn, according to Dong et al. [49], compounds formed as a result of the attack of HO radicals on the nitrogen atom of the amide group were present among the degradation products of SAD.
Figure 3 presents the two proposed identified products of SAD degradation, their molecular and structural formulas, the masses of the monoisotopic molecular ions (M+H+), and the error in mass determination (in ppm).
Both products were formed as a result of the SAD transformation, regardless of the pH, as a result of the attack by HO radicals on the aromatic ring of SAD. They were identified immediately after sampling. These compounds may undergo further changes, likely in the presence of oxygen dissolved in water. Because the -OH group is attached directly to the aromatic ring, the products of SAD degradation cannot be a competitive inhibitor of dihydrofolate synthesis; therefore, they should not exhibit antimicrobial activity and phytotoxicity.
Figure 4, Figure 5, Figure 6, Figure 7 and Figure 8 present the proposed structural and molecular formulas of the degradation products of other SNs after 120 min of irradiation in the presence of TiO2-P25. Compounds that may act as competitive inhibitors of dihydrofolate synthesis are marked in orange. Compounds that can be converted to the active form in the environment are marked in blue.
The analysis of the results (Figure 4, Figure 5, Figure 6 and Figure 7) indicated that the qualitative composition of the solutions after SN degradation did not differ, with few exceptions, in the neutral and alkaline environments. More differences were found in the acidic environment. Under these conditions, the compounds with the hydroxyl group in the aromatic and/or heterocyclic ring and those resulting from amide hydrolysis were identified as the main products. Similar compounds were also reported by other researchers [41,50,51,52,53]. These compounds were formed as a result of the attack by HO radicals in regions with higher electron density in SN molecules [54].
The compounds formed as a result of the oxidation of the amino group to -NHOH and -NO2 and those containing a degraded and oxidized heterocyclic ring were found among the products of photocatalytic degradation. More interesting is the fact that previously unknown azo compounds were also detected. These products were found in the solutions of SMX (Figure 4E,F), STZ (Figure 5E,G,H), and SMR (Figure 7G,H) and were present at pH 3. In the case of STZ, they were also detected in the alkaline environment. The azo products could derive from the recombination of organic radicals (cation radicals) generated in solutions. The generation of cation radicals is more likely in an acidic environment. Bhat and Gogate reported the mechanism of their formation as a result of the reactions initiated by O2−• radicals [54]. The azo derivatives may undergo transformation in the environment to the initial SNs.
Other products were identified in solutions after the photocatalytic degradation of SCP (Figure 8). Compounds D, E, F, and I resulted from the elimination of the sulfone group. In turn, a similar product was identified in the post-reaction SMR solution (Figure 7B). Shah and Hao presented the compounds formed during sulfamethoxypyridazine degradation as having similar structures [55]. They proposed a mechanism and the thermodynamic aspects of the HO-initiated reactions. In our opinion, an important stage in this process is the generation of the (NH2)C6H6 radical as a result of an attack by the O2−• radical and the breakage of the C-S bonds. The stable hydrogen sulfate(VI) anion can also form in this reaction.
As mentioned above, the majority of the degradation products formed in reactions of SNs with HO radicals. This was consistent with the assumption that the susceptibility of SNs to reactions with these radicals may be of key importance when considering the degradation rate.

3.3. Toxicity Prediction of the Degradation Products of SMX

Most of the products described in Section 3.2 (Figure 4, Figure 5, Figure 6, Figure 7 and Figure 8, marked in orange) contain a moiety that is responsible for the biological activity of SNs. However, it is not excluded that other photocatalytic degradation products of SNs are also hazardous to the environment. Figure 9 shows the predicted chronic toxicity of the products identified in solutions after the photocatalytic degradation of SMX and the aliphatic degradation products of SNs reported in the literature [48]. Chronic aquatic toxicity (ChV) inversion was used as a measure of toxicity to representative environmental (aquatic) organisms (unicellular and multicellular). ChV is defined as the geometric mean of the no-observed-effect concentration (NOEC) and the lowest-observed-effect concentration (LOEC).
The predicted results of the ECOSAR model indicated that substances formed during the photocatalytic degradation of SNs may be highly toxic to aquatic organisms, e.g., the products containing the nitro group (D) and the azo bond (F). However, the estimated content of toxic substances in the analyzed samples was very low (Figure 10).
The complete mineralization of antibiotics via photocatalysis is much slower than their degradation and the inactivation of antimicrobial activity, because it requires a significantly longer irradiation time and is ineffective in real wastewater [40]. For these reasons, the irradiation time necessary to achieve the inhibition of antimicrobial activity in solutions containing antibiotics residues was sufficient. We showed that the aliphatic compounds remaining in solution were likely non-toxic or little toxic to environmental aquatic organisms (Figure 9). In addition, subtoxic amounts of SNs to bacteria and their biologically active derivatives also remained in the solutions after irradiation (Figure 4, Figure 5, Figure 6, Figure 7 and Figure 8). The high phytotoxicity of these compounds to aquatic plants could be a serious threat, especially to vulnerable ecosystems. A complete assessment of the photocatalytic efficiency of SN degradation should include ecotoxicity tests with the use of plant indicators; however, the long testing time is a major disadvantage.

4. Conclusions

The pH of the solution significantly affected the photocatalytic degradation rate of SNs in the presence of TiO2 P25. This effect was particularly evident in the acidic and alkaline environments and likely depended on the susceptibility of SNs to attack by HO radicals. The pH did not affect the type of degradation products of SAD. In the case of other SNs, there were significant differences among the products identified in acidic, neutral, and alkaline solutions. In the post-reaction mixture, we identified the compounds resulting from the substitution of the aromatic rings with a hydroxyl group; the amide hydrolysis products; the hydroxylamine, azo, and nitro derivatives; and the compounds formed via the elimination of the sulfone group. The azo and nitro compounds may be potentially more toxic to aquatic ecosystems than the initial SNs. Some degradation products of SNs may exhibit marked bacteriostatic activity and high phytotoxicity. The biological activity analysis of the products is an important determinant to assess the effectiveness of the treatment of wastewater containing SN residues using the photocatalytic method. In turn, the pH change of the photocatalytic process to increase the treatment efficiency is not justified.

Author Contributions

Conceptualization, methodology, and writing—original draft preparation, E.A. and W.B.; investigation software, E.M. and D.S.; funding acquisition, data curation, and writing—review and editing, W.Z.-D. All authors have read and agreed to the published version of the manuscript.

Funding

This research study was funded by a Medical University of Silesia grant, number PCN-1-002/N/1/F.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author. The data are not publicity available due to the very large sizes of the chromatographic files.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. European Medicines Agency. Sales of Veterinary Antimicrobial Agents in 31 European Countries in 2019 and 2020. Available online: https://www.ema.europa.eu/en/documents/report/sales-veterinary-antimicrobial-agents-31-european-countries-2019-2020-trends-2010-2020-eleventh_en.pdf (accessed on 1 August 2022).
  2. FDA. Antimicrobials Sold or Distributed for Use in Food-Producing Animals. Available online: https://www.fda.gov/animal-veterinary/cvm-updates/fda-releases-annual-summary-report-antimicrobials-sold-or-distributed-2020-use-food-producing (accessed on 1 August 2022).
  3. Adamek, E.; Baran, W.; Sobczak, A. Assessment of the Biodegradability of Selected Sulfa Drugs in Two Polluted Rivers in Poland: Effects of Seasonal Variations, Accidental Contamination, Turbidity and Salinity. J. Hazard. Mater. 2016, 313, 147–158. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, H.; Chu, Y.; Fang, C. Occurrence of Veterinary Antibiotics in Swine Manure from Large-Scale Feedlots in Zhejiang Province, China. Bull. Environ. Contam Toxicol. 2017, 98, 472–477. [Google Scholar] [CrossRef] [PubMed]
  5. Winfieldunited United. Asulam 3.3. Available online: https://www.winfieldunited.com/products/herbicides/asulam33/84 (accessed on 1 August 2022).
  6. Arena, M.; Auteri, D.; Barmaz, S.; Brancato, A.; Brocca, D.; Bura, L.; Chiusolo, A.; Court Marques, D.; Crivellente, F.; de Lentdecker, C.; et al. Peer Review of the Pesticide Risk Assessment of the Active Substance Asulam (Variant Evaluated Asulam-sodium). EFSA J. 2018, 16, e05251. [Google Scholar] [CrossRef]
  7. Alvarez, F.; Arena, M.; Auteri, D.; Borroto, J.; Brancato, A.; Carrasco Cabrera, L.; Castoldi, A.F.; Chiusolo, A.; Colagiorgi, A.; Colas, M.; et al. Updated Peer Review of the Pesticide Risk Assessment of the Active Substance Asulam (Variant Evaluated Asulam-sodium). EFSA J. 2021, 19, e06921. [Google Scholar] [CrossRef]
  8. Dekker, J.; Duke, S. Herbicide-Resistant Field Crops. Adv. Agron. 1995, 54, 69–116. [Google Scholar]
  9. Duke, S.O. Herbicide and Pharmaceutical Relationships. Weed Sci. 2010, 58, 334–339. [Google Scholar] [CrossRef]
  10. Ferrari, B.; Mons, R.; Vollat, B.; Fraysse, B.; Paxéus, N.; lo Giudice, R.; Pollio, A.; Garric, J. Environmental Risk Assessment of Six Human Pharmaceuticals: Are The Current Environmental Risk Assessment Procedures Sufficient For The Protection of The Aquatic Environment? Environ. Toxicol. Chem. 2004, 23, 1344. [Google Scholar] [CrossRef] [Green Version]
  11. Białk-Bielińska, A.; Stolte, S.; Arning, J.; Uebers, U.; Böschen, A.; Stepnowski, P.; Matzke, M. Ecotoxicity Evaluation of Selected Sulfonamides. Chemosphere 2011, 85, 928–933. [Google Scholar] [CrossRef] [PubMed]
  12. Kim, Y.; Choi, K.; Jung, J.; Park, S.; Kim, P.-G.; Park, J. Aquatic Toxicity of Acetaminophen, Carbamazepine, Cimetidine, Diltiazem and Six Major Sulfonamides, and Their Potential Ecological Risks in Korea. Environ. Int. 2007, 33, 370–375. [Google Scholar] [CrossRef] [PubMed]
  13. Jung, J.; Kim, Y.; Kim, J.; Jeong, D.-H.; Choi, K. Environmental Levels of Ultraviolet Light Potentiate the Toxicity of Sulfonamide Antibiotics in Daphnia Magna. Ecotoxicology 2008, 17, 37–45. [Google Scholar] [CrossRef] [PubMed]
  14. Isidori, M.; Lavorgna, M.; Nardelli, A.; Pascarella, L.; Parrella, A. Toxic and Genotoxic Evaluation of Six Antibiotics on Non-Target Organisms. Sci. Total Environ. 2005, 346, 87–98. [Google Scholar] [CrossRef]
  15. Hillis, D.G.; Antunes, P.; Sibley, P.K.; Klironomos, J.N.; Solomon, K.R. Structural Responses of Daucus Carota Root-Organ Cultures and the Arbuscular Mycorrhizal Fungus, Glomus Intraradices, to 12 Pharmaceuticals. Chemosphere 2008, 73, 344–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Białk-Bielińska, A.; Caban, M.; Pieczyńska, A.; Stepnowski, P.; Stolte, S. Mixture Toxicity of Six Sulfonamides and Their Two Transformation Products to Green Algae Scenedesmus Vacuolatus and Duckweed Lemna Minor. Chemosphere 2017, 173, 542–550. [Google Scholar] [CrossRef] [PubMed]
  17. Thiele-Bruhn, S.; Beck, I.-C. Effects of Sulfonamide and Tetracycline Antibiotics on Soil Microbial Activity and Microbial Biomass. Chemosphere 2005, 59, 457–465. [Google Scholar] [CrossRef]
  18. Kemper, N. Veterinary Antibiotics in the Aquatic and Terrestrial Environment. Ecol. Indic. 2008, 8, 1–13. [Google Scholar] [CrossRef]
  19. Pazda, M.; Kumirska, J.; Stepnowski, P.; Mulkiewicz, E. Antibiotic Resistance Genes Identified in Wastewater Treatment Plant Systems—A Review. Sci. Total Environ. 2019, 697, 134023. [Google Scholar] [CrossRef]
  20. Długosz, M.; Żmudzki, P.; Kwiecień, A.; Szczubiałka, K.; Krzek, J.; Nowakowska, M. Photocatalytic Degradation of Sulfamethoxazole in Aqueous Solution Using a Floating TiO2-Expanded Perlite Photocatalyst. J. Hazard. Mater. 2015, 298, 146–153. [Google Scholar] [CrossRef]
  21. Biancullo, F.; Moreira, N.F.F.; Ribeiro, A.R.; Manaia, C.M.; Faria, J.L.; Nunes, O.C.; Castro-Silva, S.M.; Silva, A.M.T. Heterogeneous Photocatalysis Using UVA-LEDs for the Removal of Antibiotics and Antibiotic Resistant Bacteria from Urban Wastewater Treatment Plant Effluents. Chem. Eng. J. 2019, 367, 304–313. [Google Scholar] [CrossRef]
  22. Park, Y.; Kim, S.; Kim, J.; Khan, S.; Han, C. UV/TiO2 Photocatalysis as an Efficient Livestock Wastewater Quaternary Treatment for Antibiotics Removal. Water 2022, 14, 958. [Google Scholar] [CrossRef]
  23. Adamek, E.; Baran, W.; Ziemiańska, J.; Sobczak, A. Effect of FeCl3 on Sulfonamide Removal and Reduction of Antimicrobial Activity of Wastewater in a Photocatalytic Process with TiO2. Appl. Catal. B 2012, 126, 29–38. [Google Scholar] [CrossRef]
  24. Zhang, L.; Ma, P.; Dai, L.; Li, S.; Yu, W.; Guan, J. In Situ Crystallization and Growth of TiO2 Nanospheres between MXene Layers for Improved Adsorption and Visible Light Photocatalysis. Catal. Sci. Technol. 2021, 11, 3834–3844. [Google Scholar] [CrossRef]
  25. Liu, X.; Wu, F.; Deng, N. Photoproduction of Hydroxyl Radicals in Aqueous Solution with Algae under High-Pressure Mercury Lamp. Environ. Sci. Technol. 2004, 38, 296–299. [Google Scholar] [CrossRef] [PubMed]
  26. Joseph, J.M.; Varghese, R.; Aravindakumar, C.T. Photoproduction of Hydroxyl Radicals from Fe(III)-Hydroxy Complex: A Quantitative Assessment. J. Photochem. Photobiol. A Chem. 2001, 146, 67–73. [Google Scholar] [CrossRef]
  27. Baran, W.; Adamek, E.; Sobczak, A.; Makowski, A. Photocatalytic Degradation of Sulfa Drugs with TiO2, Fe Salts and TiO2/FeCl3 in Aquatic Environment—Kinetics and Degradation Pathway. Appl. Catal. B 2009, 90, 516–525. [Google Scholar] [CrossRef]
  28. Jiang, L.; Zhou, S.; Yang, J.; Wang, H.; Yu, H.; Chen, H.; Zhao, Y.; Yuan, X.; Chu, W.; Li, H. Near-Infrared Light Responsive TiO2 for Efficient Solar Energy Utilization. Adv. Funct. Mater. 2022, 32, 2108977. [Google Scholar] [CrossRef]
  29. Jiang, L.; Yang, J.; Zhou, S.; Yu, H.; Liang, J.; Chu, W.; Li, H.; Wang, H.; Wu, Z.; Yuan, X. Strategies to Extend Near-Infrared Light Harvest of Polymer Carbon Nitride Photocatalysts. Coord. Chem. Rev. 2021, 439, 213947. [Google Scholar] [CrossRef]
  30. Sági, G.; Csay, T.; Szabó, L.; Pátzay, G.; Csonka, E.; Takács, E.; Wojnárovits, L. Analytical Approaches to the OH Radical Induced Degradation of Sulfonamide Antibiotics in Dilute Aqueous Solutions. J. Pharm. Biomed. Anal. 2015, 106, 52–60. [Google Scholar] [CrossRef]
  31. Xiang, Q.; Yu, J.; Wong, P.K. Quantitative Characterization of Hydroxyl Radicals Produced by Various Photocatalysts. J. Colloid. Interface Sci. 2011, 357, 163–167. [Google Scholar] [CrossRef]
  32. Shinde, S.S.; Bhosale, C.H.; Rajpure, K.Y. Kinetic Analysis of Heterogeneous Photocatalysis: Role of Hydroxyl Radicals. Catal. Rev. 2013, 55, 79–133. [Google Scholar] [CrossRef]
  33. Ahmed, S.N.; Haider, W. Heterogeneous Photocatalysis and Its Potential Applications in Water and Wastewater Treatment: A Review. Nanotechnology 2018, 29, 342001. [Google Scholar] [CrossRef] [Green Version]
  34. Náfrádi, M.; Veréb, G.; Firak, D.S.; Alapi, T. Photocatalysis: Introduction, Mechanism, and Effective Parameters. In Green Photocatalytic Semiconductors; Springer: Cham, Switzerland, 2022; pp. 3–31. [Google Scholar]
  35. Baran, W.; Cholewiński, M.; Sobczak, A.; Adamek, E. A New Mechanism of the Selective Photodegradation of Antibiotics in the Catalytic System Containing TiO2 and the Inorganic Cations. Int. J. Mol. Sci. 2021, 22, 8696. [Google Scholar] [CrossRef] [PubMed]
  36. Islam Molla, M.A.; Tateishi, I.; Furukawa, M.; Katsumata, H.; Suzuki, T.; Kaneco, S. Evaluation of Reaction Mechanism for Photocatalytic Degradation of Dye with Self-Sensitized TiO2 under Visible Light Irradiation. Open J. Inorg. Non-Met. Mater. 2017, 7, 1–7. [Google Scholar] [CrossRef] [Green Version]
  37. Mezyk, S.P.; Neubauer, T.J.; Cooper, W.J.; Peller, J.R. Free-Radical-Induced Oxidative and Reductive Degradation of Sulfa Drugs in Water: Absolute Kinetics and Efficiencies of Hydroxyl Radical and Hydrated Electron Reactions. J. Phys. Chem. A 2007, 111, 9019–9024. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, J.; Nosaka, Y. Mechanism of the OH Radical Generation in Photocatalysis with TiO2 of Different Crystalline Types. J. Phys. Chem. C 2014, 118, 10824–10832. [Google Scholar] [CrossRef]
  39. Guo, C.; Wang, K.; Hou, S.; Wan, L.; Lv, J.; Zhang, Y.; Qu, X.; Chen, S.; Xu, J. H2O2 and/or TiO2 Photocatalysis under UV Irradiation for the Removal of Antibiotic Resistant Bacteria and Their Antibiotic Resistance Genes. J. Hazard. Mater. 2017, 323, 710–718. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Adamek, E.; Baran, W.; Sobczak, A. Photocatalytic Degradation of Veterinary Antibiotics: Biodegradability and Antimicrobial Activity of Intermediates. Process. Saf. Environ. Prot. 2016, 103, 1–9. [Google Scholar] [CrossRef]
  41. Yang, H.; Li, G.; An, T.; Gao, Y.; Fu, J. Photocatalytic Degradation Kinetics and Mechanism of Environmental Pharmaceuticals in Aqueous Suspension of TiO2: A Case of Sulfa Drugs. Catal. Today 2010, 153, 200–207. [Google Scholar] [CrossRef]
  42. Adamek, E.; Baran, W.; Sobczak, A. Effect of FeCl3 on the Photocatalytic Processes Initiated by UVa and Vis Light in the Presence of TiO2–P25. Appl. Catal. B 2015, 172–173, 136–144. [Google Scholar] [CrossRef]
  43. Nakabayashi, Y.; Nosaka, Y. OH Radical Formation at Distinct Faces of Rutile TiO2 Crystal in the Procedure of Photoelectrochemical Water Oxidation. J. Phys. Chem. C 2013, 117, 23832–23839. [Google Scholar] [CrossRef]
  44. Nakabayashi, Y.; Nosaka, Y. The PH Dependence of OH Radical Formation in Photo-Electrochemical Water Oxidation with Rutile TiO2 Single Crystals. Phys. Chem. Chem. Phys. 2015, 17, 30570–30576. [Google Scholar] [CrossRef]
  45. Boreen, A.L.; Arnold, W.A.; McNeill, K. Triplet-Sensitized Photodegradation of Sulfa Drugs Containing Six-Membered Heterocyclic Groups: Identification of an SO2 Extrusion Photoproduct. Environ. Sci. Technol. 2005, 39, 3630–3638. [Google Scholar] [CrossRef] [PubMed]
  46. Zarfl, C.; Matthies, M.; Klasmeier, J. A Mechanistical Model for the Uptake of Sulfonamides by Bacteria. Chemosphere 2008, 70, 753–760. [Google Scholar] [CrossRef] [PubMed]
  47. Uhlemann, T.; Seidel, S.; Müller, C.W. Site-Specific Binding of a Water Molecule to the Sulfa Drugs Sulfamethoxazole and Sulfisoxazole: A Laser-Desorption Isomer-Specific UV and IR Study. Phys. Chem. Chem. Phys. 2018, 20, 6891–6904. [Google Scholar] [CrossRef] [PubMed]
  48. Tačić, A.; Nikolić, V.; Nikolić, L.; Savić, I. Antimicrobial Sulfonamide Drug. Adv. Technol. 2017, 6, 58–71. [Google Scholar] [CrossRef] [Green Version]
  49. Dong, S.; Pi, Y.; Li, Q.; Hu, L.; Li, Y.; Han, X.; Wang, J.; Sun, J. Solar Photocatalytic Degradation of Sulfanilamide by BiOCl/Reduced Graphene Oxide Nanocomposites: Mechanism and Degradation Pathways. J. Alloys Compd. 2016, 663, 1–9. [Google Scholar] [CrossRef]
  50. Pang, R.; Li, N.; Hou, Z.; Huang, J.; Yue, C.; Cai, Y.; Song, J. Degradation of Sulfonamide Antibiotics and a Structurally Related Compound by Chlorine Dioxide: Efficiency, Kinetics, Potential Products and Pathways. Chem. Eng. J. 2023, 451, 138502. [Google Scholar] [CrossRef]
  51. Zhang, K.; Luo, Z.; Zhang, T.; Gao, N.; Ma, Y. Degradation Effect of Sulfa Antibiotics by Potassium Ferrate Combined with Ultrasound (Fe(VI)-US). Biomed Res. Int. 2015, 2015, 1–12. [Google Scholar] [CrossRef] [Green Version]
  52. Calza, P.; Medana, C.; Pazzi, M.; Baiocchi, C.; Pelizzetti, E. Photocatalytic Transformations of Sulphonamides on Titanium Dioxide. Appl. Catal. B 2004, 53, 63–69. [Google Scholar] [CrossRef]
  53. García-Galán, M.J.; Silvia Díaz-Cruz, M.; Barceló, D. Identification and Determination of Metabolites and Degradation Products of Sulfonamide Antibiotics. TrAC Trends Anal. Chem. 2008, 27, 1008–1022. [Google Scholar] [CrossRef]
  54. Bhat, A.P.; Gogate, P.R. Degradation of Nitrogen-Containing Hazardous Compounds Using Advanced Oxidation Processes: A Review on Aliphatic and Aromatic Amines, Dyes, and Pesticides. J. Hazard. Mater. 2021, 403, 123657. [Google Scholar] [CrossRef]
  55. Shah, S.; Hao, C. Quantum Chemical Investigation on Photodegradation Mechanisms of Sulfamethoxypyridazine with Dissolved Inorganic Matter and Hydroxyl Radical. J. Environ. Sci. 2017, 57, 85–92. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Reaction rate constants of the photocatalytic degradation of SNs in acidic, neutral, and alkaline environments.
Figure 1. Reaction rate constants of the photocatalytic degradation of SNs in acidic, neutral, and alkaline environments.
Toxics 10 00655 g001
Figure 2. Relationship between pKa2 and the degradation rate constant of SNs at pH 10.
Figure 2. Relationship between pKa2 and the degradation rate constant of SNs at pH 10.
Toxics 10 00655 g002
Figure 3. Products of the photocatalytic degradation of SAD after 120 min of UV irradiation in the presence of TiO2-P25. (*) primary amide hydrolysis occurs during QTof analysis.
Figure 3. Products of the photocatalytic degradation of SAD after 120 min of UV irradiation in the presence of TiO2-P25. (*) primary amide hydrolysis occurs during QTof analysis.
Toxics 10 00655 g003
Figure 4. Products of the photocatalytic degradation of SMX after 120 min of UV irradiation in the presence of TiO2-P25. (*) primary amide hydrolysis occurs during QTof analysis.
Figure 4. Products of the photocatalytic degradation of SMX after 120 min of UV irradiation in the presence of TiO2-P25. (*) primary amide hydrolysis occurs during QTof analysis.
Toxics 10 00655 g004
Figure 5. Products of the photocatalytic degradation of STZ after 120 min of UV irradiation in the presence of TiO2-P25. (*) primary amide hydrolysis occurs during QTof analysis.
Figure 5. Products of the photocatalytic degradation of STZ after 120 min of UV irradiation in the presence of TiO2-P25. (*) primary amide hydrolysis occurs during QTof analysis.
Toxics 10 00655 g005
Figure 6. Products of the photocatalytic degradation of SFF after 120 min of UV irradiation in the presence of TiO2-P25. (*) primary amide hydrolysis occurs during QTof analysis.
Figure 6. Products of the photocatalytic degradation of SFF after 120 min of UV irradiation in the presence of TiO2-P25. (*) primary amide hydrolysis occurs during QTof analysis.
Toxics 10 00655 g006
Figure 7. Products of the photocatalytic degradation of SMR after 120 min of UV irradiation in the presence of TiO2-P25. (*) primary amide hydrolysis occurs during QTof analysis.
Figure 7. Products of the photocatalytic degradation of SMR after 120 min of UV irradiation in the presence of TiO2-P25. (*) primary amide hydrolysis occurs during QTof analysis.
Toxics 10 00655 g007
Figure 8. Products of the photocatalytic degradation of SCP after 120 min of UV irradiation in the presence of TiO2-P25.
Figure 8. Products of the photocatalytic degradation of SCP after 120 min of UV irradiation in the presence of TiO2-P25.
Toxics 10 00655 g008
Figure 9. Predicted chronic toxicity of SMX, of its photocatalytic degradation products (compounds marked with letters A through I in Figure 4), and of maleic (M…), fumaric (Fu…), acetic (Ac…), oxylatic (Ox…), and formic (Fo…) acids to selected aquatic organisms using the ECOSAR model.
Figure 9. Predicted chronic toxicity of SMX, of its photocatalytic degradation products (compounds marked with letters A through I in Figure 4), and of maleic (M…), fumaric (Fu…), acetic (Ac…), oxylatic (Ox…), and formic (Fo…) acids to selected aquatic organisms using the ECOSAR model.
Toxics 10 00655 g009
Figure 10. Chromatogram of SMX solution after 120 min of irradiation in the presence of TiO2 P25 at pH 3.
Figure 10. Chromatogram of SMX solution after 120 min of irradiation in the presence of TiO2 P25 at pH 3.
Toxics 10 00655 g010
Table 1. Characteristics of the studied SNs.
Table 1. Characteristics of the studied SNs.
Name/CASAbbreviationStructural FormulaPurity/ManufacturerpKa1 apKa2 a
Sulfanilamide
63-74-1
SADToxics 10 00655 i00198.0%;
Fluka
6.92 ± 0.1410.35 ± 0.21
Sulfamethoxazole
723-46-6
SMXToxics 10 00655 i00299.0%;
Fluka
1.86 ± 0.325.73 ± 0.20
Sulfathiazole
72-14-0
STZToxics 10 00655 i00399.0%; Sigma-Aldrich2.22 ± 0.277.15 ± 0.12
Sulfisoxazole
127-69-5
SFFToxics 10 00655 i00499.0%; Sigma-Aldrich1.69 ± 0.275.0 ± 0.0
Sulfamerazine
127-79-7
SMRToxics 10 00655 i00599.0%; Sigma-Aldrich2.21 ± 0.156.92 ± 0.14
Sulfachloro-pyridazine
80-32-0
SCPToxics 10 00655 i00698.0%; Sigma-Aldrich2 ± 3 b5.90 ± 0.30 b
(a) [42]; (b) [45].
Table 2. Chromatographic separation conditions.
Table 2. Chromatographic separation conditions.
Eluent Gradient
Time (min)H2O with 0.01% HCOOHCH3CN with 0.01% HCOOHFlow Rate (mL/min)
For SAD
099.9%0.1%0.30
299.9%0.1%0.30
2.599.0%1.0%0.30
380.0%20%0.40
3.580.0%20%0.40
499.9%0.1%0.30
For other SNs
095%5%0.35
3.060%40%
3.320%80%
3.595%5%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sapińska, D.; Adamek, E.; Masternak, E.; Zielińska-Danch, W.; Baran, W. Influence of pH on the Kinetics and Products of Photocatalytic Degradation of Sulfonamides in Aqueous Solutions. Toxics 2022, 10, 655. https://doi.org/10.3390/toxics10110655

AMA Style

Sapińska D, Adamek E, Masternak E, Zielińska-Danch W, Baran W. Influence of pH on the Kinetics and Products of Photocatalytic Degradation of Sulfonamides in Aqueous Solutions. Toxics. 2022; 10(11):655. https://doi.org/10.3390/toxics10110655

Chicago/Turabian Style

Sapińska, Dominika, Ewa Adamek, Ewa Masternak, Wioleta Zielińska-Danch, and Wojciech Baran. 2022. "Influence of pH on the Kinetics and Products of Photocatalytic Degradation of Sulfonamides in Aqueous Solutions" Toxics 10, no. 11: 655. https://doi.org/10.3390/toxics10110655

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