Next Article in Journal
Formation of Amphiphilic Molecules from the Most Common Marine Polysaccharides, toward a Sustainable Alternative?
Previous Article in Journal
Inhibitory Effects of Cucurbitane-Type Triterpenoids from Momordica charantia Fruit on Lipopolysaccharide-Stimulated Pro-Inflammatory Cytokine Production in Bone Marrow-Derived Dendritic Cells
Previous Article in Special Issue
Synthesis of 4-Arylselanyl-1H-1,2,3-triazoles from Selenium-Containing Carbinols
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Antioxidant Activity of Bis-Aniline-Derived Diselenides as GPx Mimics

by
Giancarlo V. Botteselle
1,*,
Welman C. Elias
2,
Luana Bettanin
2,
Rômulo F. S. Canto
3,
Drielly N. O. Salin
2,
Flavio A. R. Barbosa
2,
Sumbal Saba
4,
Hugo Gallardo
2,
Gianluca Ciancaleoni
5,
Josiel B. Domingos
2,
Jamal Rafique
6,* and
Antonio L. Braga
2,7,*
1
Departamento de Química, Universidade Estadual do Centro-Oeste (UNICENTRO), Guarapuava 85040-167, PR, Brazil
2
Departamento de Química, Universidade Federal de Santa Catarina (UFSC), Florianópolis 88040-970, SC, Brazil
3
Programa de Pós-Graduação em Ciências da Saúde, Universidade Federal de Ciências da Saúde de Porto Alegre (UFCSPA), Porto Alegre 90050-170, RS, Brazil
4
Instituto de Química—IQ, Universidade Federal de Goiás—(UFG), Goiânia 74690-900, GO, Brazil
5
Department of Chemistry and Industrial Chemistry, University of Pisa, Via G. Moruzzi 13, I-56124 Pisa, Italy
6
Instituto de Química—INQUI, Universidade Federal do Mato Grosso do Sul (UFMS), Campo Grande 79074-460, MS, Brazil
7
Department of Chemical Sciences, Faculty of Science, University of Johannesburg, Doornfontein 2028, South Africa
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(15), 4446; https://doi.org/10.3390/molecules26154446
Submission received: 23 June 2021 / Revised: 14 July 2021 / Accepted: 20 July 2021 / Published: 23 July 2021
(This article belongs to the Special Issue Young Researchers in Sulfur/Selenium Chemistry)

Abstract

:
Herein, we describe a simple and efficient route to access aniline-derived diselenides and evaluate their antioxidant/GPx-mimetic properties. The diselenides were obtained in good yields via ipso-substitution/reduction from the readily available 2-nitroaromatic halides (Cl, Br, I). These diselenides present GPx-mimetic properties, showing better antioxidant activity than the standard GPx-mimetic compounds, ebselen and diphenyl diselenide. DFT analysis demonstrated that the electronic properties of the substituents determine the charge delocalization and the partial charge on selenium, which correlate with the catalytic performances. The amino group concurs in the stabilization of the selenolate intermediate through a hydrogen bond with the selenium.

1. Introduction

In recent years, there has been an increasing interest in synthetic organoselenium compounds, mainly due to their properties as synthetic intermediates in organic transformations [1,2,3] and material sciences [4,5], as well as in medicinal chemistry [6,7,8,9]. These compounds have been recently described as good antioxidants [10,11], also presenting anti-inflammatory [12,13], antibacterial [14], antiviral [15], anticancer [16,17,18,19], anti-Alzheimer’s [20,21,22,23] and other activities [24,25,26,27,28]. Furthermore, in relation to the current pandemic of COVID-19, there are some interesting studies available that demonstrate the effectiveness of organoselenium compound (Ebselen) as an antiviral molecule, Figure 1 [29,30,31].
Among organoselenium compounds, diorganyl diselenides present important antioxidant and anticancer properties mainly because of the ability of these diselenides to act as mimetics of the enzyme glutathione peroxidase (GPx) [6,8,11]. This selenoenzyme possesses a residue of selenocysteine in its active site and is responsible for the reduction of peroxides to water in our organism, protecting it from oxidative stress and related diseases [32].
Notable among the polyfunctionalized diselenides, the presence of an amino or carbonyl group in close proximity to the selenium moiety has some unique biological features due to non-bonding interactions [33,34,35]. For example, the bis-2-aniline diselenide is reported to be a good antioxidant, preventing the oxidative stress caused by peroxynitrite and hydroperoxides [36,37,38]. Moreover, the aniline-derived diselenides, mainly with an amino group in the ortho position, give these compounds two possible reactive centers, Se-Se bond cleavage and the unshared pair of electrons on the nitrogen. This makes this class of compounds extremely flexible in functional group interconversions, making it appropriate for several transformations, mainly in the formation of selenium-containing heterocycles, such as selenamides [39], benzoselenazines [40], benzoselenazoles [41] and triazole diselenides [42].
Recently, we reported a new robust methodology for the synthesis of o-aniline-derived diselenides from the reduction of o-nitrobenzene diselenides. As part of our wider research program aimed at efficient methodologies for the synthesis of organoselenium compounds and their biological evaluation [43,44,45,46,47,48,49,50], herein, we report the application of o-aniline-derived diselenides as potential GPx mimics. For this purpose, different physical-chemical studies were performed to demonstrate their biological properties, i.e., kinetic profile. Furthermore, DFT studies were also carried out in order to study the electronic properties of the substituents for determining the charge delocalization on the selenium atom and its influence on catalytic performance.

2. Results and Discussion

The bis-o-nitrobenzene diselenides were initially prepared through the nucleophilic aromatic substitution of o-halonitrobenzenes with K2Se2 (generated in situ) from a modified simple methodology [51]. After this, we carried out the reduction of bis-o-nitrobenzene diselenides from a well-established procedure described in the literature [52] using low-cost iron sulfate heptahydrate (FeSO4.7H2O) to synthesize the aniline-derived diselenides (Scheme 1).
The aniline-derived diselenides 3a–e were evaluated with regard to GPx-like antioxidant activity. The catalytic parameters were obtained using the Tomoda [53] reaction model, where the synthesized diselenides were applied as catalysts in the formation of diphenyl disulfide (PhSSPh) through the reduction of hydrogen peroxide (H2O2) in the presence of thiophenol (PhSH), which is accompanied by an increase in UV/vis absorbance in 305 nm (Figure 2a). Then, the absorbance was plotted against diphenyl disulfide concentration to determine the molar absorptivity at 305 nm (Figure 2b).
The catalytic parameters were obtained by fitting the kinetic profiles, that is, initial rate versus initial PhSH concentration, as shown in Figure 3 for diselenide 3b (for other compounds, see Figures S31–S38, Pg. S22–S25 in Supplementary Materials), with the Michaelis–Menten equation (Equation (1)).
Table 1 shows the catalytic constant (kcat), the Michaelis–Menten constant (Km) and the catalytic efficiency (η where η = kcat/Km) for the reaction with the aniline-derived diselenides 3a to 3e and, for comparison, the well-known catalysts ebselen [35] and diphenyl diselenide [53].
The results show that the catalytic efficiency of the aniline-derived diselenides is structure-dependent, especially regarding the electronic character of the substituents at the para position related to selenium, with an increase in the catalytic efficiency with the electron-withdrawal capacity of the substituent (compound 3b), once cleavage of the Se-Se bond is facilitated. These results suggest that the mechanism of these catalyzed reactions involves the formation of a zwitterionic form of the selenolate intermediate with a negative density charge in the selenium atom (Scheme 2a), which is similar to the mechanism proposed by Tomoda et al. [53].
The structures 3ae have been optimized by Density Functional Theory (DFT) at the BP86-D3/def2-TZVP level of theory, using the zero-order regular approximation (ZORA) to take the relativistic effects into account. In all the optimized geometries (except 3b), an intramolecular hydrogen bond (HB) exists between the two amine moieties, with distances that range from 2.624 (3d) to 3.139 (3a) Å. In the case of 3b, the electron-withdrawing -CF3 group likely makes the lone pair of the nitrogen less available for HBs. The electronic effect of the group in the para position influences all the atomic charges of the diselenide system. Indeed, the atomic charges have been computed through the Natural Population Analysis (see Computational Details) as implemented in NBO 6.0, and for the selenium, it ranges from 0.063 to 0.123 e for 3c (the most electron-donating group) and 3b (the most electron-withdrawing one), respectively. The atomic charge on the selenium qualitatively correlates with η, according to which the best catalysts have a more positive charge on the selenium and a less negative charge on the nitrogen (Figure 4 and Supplementary Materials). In addition, a similar correlation can be observed between the atomic charge of the ammonium-selenolate and η: in this case, the best catalysts have a less negative charge on the selenium, leading to a larger degree of charge delocalization and, consequently, a more stable intermediate. This is in agreement with the mechanism proposed by Tomoda [53]. Furthermore, the hydrogen bonding between the ammonium protons and the selenolate moiety is quite strong and stabilizes the intermediate, having an orbital interaction of 8–9 kcal/mol depending on the substituent (Supplementary Materials), hence making the catalyst more active.
Tomoda et al. [53] also proposed that another reactive intermediate is formed in the initial step from the reaction of the diselenide with PhSH, that is, the selenyl sulfide (Scheme 2b). In the case of the aniline-derived diselenides, it seems that the formation of this intermediate is destabilized by the inductive electron donor capacity of the amine groups at the ortho position, reflected in their lower catalytic efficiency when compared with the diphenyl diselenide (Table 1, entries 3 and 2, respectively).
Of the aniline-derived diselenides, the highest catalytic efficiency was observed for compound 3b (Table 1, entry 4), which was 5 and 2 times more active than the standards ebselen and diphenyl diselenide, respectively. It is worth noting that the diselenide 3b was more effective than ebselen, which is a pre-clinical drug candidate with pronounced biological activities, including, recently, the inhibition of protease Mpro from COVID-19 (SARS-CoV-2) virus [29,30,31].
Due to the high antioxidant activity of the diselenide 3b as a mimetic of GPx, we decided to investigate the effectiveness of this methodology at the gram scale. Thus, we performed the reaction from 20.0 mmol (5.40 g) of the o-halonitrobenzene 1c to afford the desired nitro-diselenide 2b, followed by its reduction to obtain the bis-aniline-derived 3b without a significant decrease in the yields (Scheme 3), proving that this protocol could be used as a robust method in the larger-scale synthesis of this privileged structure.

3. Materials and Methods

3.1. GPx-Like Experimental Procedure

The kinetic profile of the oxidation reaction was conducted in a UV-vis Spectrophotometer, following the wavelength of diphenyl disulfide formation at 305 nm. Spectroscopic methanol was used as solvent in the oxidation reaction, and the final volume of cuvettes was kept at 2000 µL. The H2O2 and catalyst concentration were fixed in 15 × 10−3 mol L−1 and 1 × 10−5 mol L−1 respectively, and the PhSH concentration was varied from 0.5 × 10−3 to 15 × 10−3 mol L−1. The temperature was kept at 25 °C, and each experiment was run at least 2 times.

3.2. Michaelis–Menten Equation

The GPx-like kinetic profiles were treated using the Michaelis–Menten nonlinear Equation (1):
V = k c a t [ c a t ] [ P h S H ] ( K m + [ P h S H ] )
where the V0 was the initial velocity and kcat and Km were the catalytic rate constant and Michaelis–Menten constant, respectively. The [cat] and [PhSH] represent the concentration of the catalyst and thiophenol, respectively.

3.3. Computational Details

All geometries were optimized with ORCA 4.1.0, [54] using the BP86 functional in conjunction with a triple-ζ quality basis set (ZORA-TZVP) and def2/J auxiliary basis. For heavy elements (such as selenium and bromine), relativistic effects have been accounted by using the Zeroth Order Regular Approximation (ZORA) scalar correction. The dispersion corrections were introduced using the Grimme D3-parametrized correction and the Becke−Johnson damping to the DFT energy [55]. All the diselenide structures were confirmed to be local energy minima (no imaginary frequencies). Selenolate species show an unavoidable imaginary frequency correlated with the rotation of the -NH3 moiety. The atomic charges have been computed by the Natural Population Analysis (NPA) as implemented in NBO6 [56].

4. Conclusions

In conclusion, we have developed a short and robust synthetic route for the synthesis of nitro aryl and aniline-derived diselenides in good overall yields. The aniline-derived diselenides were evaluated as GPx mimetics and the diselenide 3b substituted with the CF3 group showed the best results, being 5 and 2 times more effective as a GPx mimetic than the standard catalysts ebselen and diphenyl diselenide, respectively. Furthermore, DFT analysis was performed for all the diselenides, which demonstrated non-bonding interaction. This correlates with the GPx activities of these diselenides.

Supplementary Materials

The following are available online, 1H, and 13C NMR spectra of the synthesized compounds (3ae). Figure S1: 1H NMR (200 MHz, CDCl3) Spectrum of compound 2a. Figure S2: 13C NMR (50 MHz, CDCl3) Spectrum of compound 2a. Figure S3: HRMS spectrum of compound 2a. Figure S4: 1H NMR (200 MHz, CDCl3) Spectrum of compound 2b. Figure S5: 13C NMR (50 MHz, CDCl3) Spectrum of compound 2b. Figure S6: HRMS spectrum of compound 2b. Figure S7: 1H NMR (200 MHz, CDCl3) Spectrum of compound 2c. Figure S8: 13C NMR (50 MHz, CDCl3) Spectrum of compound 2c. Figure S9: HRMS spectrum of compound 2c. Figure S10: 1H NMR (200 MHz, CDCl3) Spectrum of compound 2d. Figure S11: 13C NMR (50 MHz, CDCl3) Spectrum of compound 2d. Figure S12: HRMS spectrum of compound 2d. Figure S13: 1H NMR (200 MHz, CDCl3) Spectrum of compound 2e. Figure S14: 13C NMR (50 MHz, CDCl3) Spectrum of compound 2e. Figure S15: ESI-MS spectrum of compound 2e. Figure S16:1H NMR (200 MHz, CDCl3) Spectrum of compound 3a. Figure S17: 13C NMR (50 MHz, CDCl3) Spectrum of compound 3a. Figure S18: 1H NMR (200 MHz, CDCl3) Spectrum of compound 3b. Figure S19: 13C NMR (50 MHz, CDCl3) Spectrum of compound 3b. Figure S20: HRMS spectrum of compound 3b. Figure S21: 1H NMR (200 MHz, CDCl3) Spectrum of compound 3c. Figure S22: 13C NMR (50 MHz, CDCl3) Spectrum of compound 3c. Figure S23: HRMS spectrum of compound 3c. Figure S24: 1H NMR (200 MHz, CDCl3) Spectrum of compound 3d. Figure S25: 13C NMR (50 MHz, CDCl3) Spectrum of compound 3d. Figure S26: HRMS spectrum of compound 3d. Figure S27: 1H NMR (200 MHz, CDCl3) Spectrum of compound 3e. Figure S28: 13C NMR (50 MHz, CDCl3) Spectrum of compound 3e. Figure S29: ESI-MS spectrum of compound 3e. Figure S30: 77Se NMR (76 MHz, CDCl3) Spectrum of compound 3b. Figure S31: Absorbance plotted against diphenyl disulfide concentration. The red line represents the linear fit. The coefficient of molar absorptivity in 305 nm was 1415 L mol−1 cm−1 (R2 = 0.9996). Figure S32: Initial rate (V0) plotted against substrate concentration. The initial rates were calculated from at least two experiments for each concentration of PhSH. The concentrations of ebselen and H2O2 were fixed at 5 × 10−5 and 15 × 10−3 mol L−1, respectively. The red line represents the Michaelis-Menten fit. Figure S33: Initial rate (V0) plotted against substrate concentration. The initial rates were calculated from at least two experiments for each concentration of PhSH. The concentrations of diphenyl disulfide and H2O2 were fixed at 5 × 10−5 and 15 × 10−3 mol L−1, respectively. The red line represents the Michaelis-Menten fit. Figure S34: Initial rate (V0) plotted against substrate concentration. The initial rates were calculated from at least two experiments for each concentration of PhSH. The concentrations of 3c and H2O2 were fixed at 5 × 10−5 and 15 × 10−3 mol L−1, respectively. The red line represents the Michaelis-Menten fit. Figure S35: Initial rate (V0) plotted against substrate concentration. The initial rates were calculated from at least two experiments for each concentration of PhSH. The concentrations of 3a and H2O2 were fixed at 5 × 10−5 and 15 × 10−3 mol L−1, respectively. The red line represents the Michaelis-Menten fit. Figure S36: Initial rate (V0) plotted against substrate concentration. The initial rates were calculated from at least two experiments for each concentration of PhSH. The concentrations of 3b and H2O2 were fixed at 5 × 10−5 and 15 × 10−3 mol L−1, respectively. The red line represents the Michaelis-Menten fit. Figure S37: Initial rate (V0) plotted against substrate concentration. The initial rates were calculated from at least two experiments for each concentration of PhSH. The concentrations of 3d and H2O2 were fixed at 5 × 10−5 and 15 × 10−3 mol L−1, respectively. The red line represents the Michaelis-Menten fit. Figure S38: Initial rate (V0) plotted against substrate concentration. The initial rates were calculated from at least two experiments for each concentration of PhSH. The concentrations of 3e and H2O2 were fixed at 5 × 10−5 and 15 × 10−3 mol L−1, respectively. The red line represents the Michaelis-Menten fit. Table S1: Atomic charges of nitrogen and selenium according to NPA. Table S2: Donor-acceptor second order perturbation analysis. Table S3: DFT-computed energies for optimized geometries (in kcal/mol).

Author Contributions

Conceptualization, G.V.B., J.R. and A.L.B.; synthesis, spectral analysis, characterizations, and reagents/materials, G.V.B., L.B., R.F.S.C., F.A.R.B., S.S., H.G. and J.R.; GPX studies, W.C.E., D.N.O.S. and J.B.D.; DFT analysis, G.C.; writing—original draft, G.V.B. and J.R. writing—review and editing, G.V.B., J.R. and A.L.B. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We gratefully acknowledge “Coordenação de Aperfeiçoamento de Pessoal de Nível Superior—CAPES” (Finance Code 001), “Conselho Nacional de Desenvolvimento Científico e Tecnológico—CNPq”, “Instituto Nacional de Ciência e Tecnologia de Catálise em Sistemas Moleculares e Nanoestruturados—INCT-Catálise/CNPq/FAPESC”, “Centro de Excelência para Pesquisa em Química Sustentável—CERSusChem” (grant 2014/50249-8), “Fundação de Amparo à Pesquisa do Estado de São Paulo—FAPESP”, (grant 2014/50249-8) “GlaxoSmithKline—GSK” (grant 2014/50249-8). G.V.B. would like to acknowledge CNPq (429831/2018-8). J.R. would like to acknowledge CNPq (433896/2018-3 and 315399/2020-1). The authors acknowledge “Laboratório Central de Biologia Molecular Estrutural (CEBIME)” (UFSC-Brazil) for the HRMS analysis.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds can be checked with the author, G.V.B.

References

  1. Shao, L.; Li, Y.; Lu, J.; Jiang, X. Recent progress in selenium-catalyzed organic reactions. Org. Chem. Front. 2019, 6, 2999–3041. [Google Scholar] [CrossRef]
  2. Arora, A.; Sinhg, S.; Oswal, P.; Nautiyal, D.; Rao, G.K.; Kumar, S.; Kumar, A. Preformed molecular complexes of metals with organoselenium ligands: Syntheses and applications in catalysis. Coord. Chem. Rev. 2021, 438, 213885. [Google Scholar] [CrossRef]
  3. Rafique, J.; Rampoin, D.S.; Azeredo, J.B.; Coelho, F.L.; Schneider, P.H.; Braga, A.L. Light-mediated seleno-functionalization of organic molecules: Recent advances. Chem. Rec. 2021. [Google Scholar] [CrossRef] [PubMed]
  4. Li, Q.; Zhang, Y.; Chen, Z.; Pan, X.; Zhan, Z.; Zhu, J.; Zhu, X. Organoselenium chemistry-based polymer synthesis. Org. Chem. Front. 2020, 7, 2815–2841. [Google Scholar] [CrossRef]
  5. Xiao, X.; Shao, Z.; Yu, L. A perspective of the engineering applications of carbon-based selenium-containing materials. Chin. Chem. Lett. 2021. [Google Scholar] [CrossRef]
  6. Nogueria, C.W.; Barbos, N.V.; Rocha, J.B.T. Toxicology and pharmacology of synthetic organoselenium compounds: An update. Arch. Toxicol. 2021, 95, 1179–1226. [Google Scholar] [CrossRef] [PubMed]
  7. Braga, A.L.; Rafique, J. Synthesis of biologically relevant small molecules containing selenium. Part A. Antioxidant compounds. In The Chemistry of Organic Selenium and Tellurium Compounds; Rappoport, Z., Ed.; Wiley: Chichester, UK, 2013; Volume 4, pp. 989–1052. [Google Scholar]
  8. Barbosa, N.V.; Nogueira, C.W.; Nogara, P.A.; Bem, A.F.; Aschner, M.; Rocha, J.B. Organoselenium compounds as mimics of selenoproteins and thiol modifier agentes. Metallomics 2017, 9, 1703–1734. [Google Scholar] [CrossRef]
  9. Braga, A.L.; Rafique, J. Synthesis of biologically relevant small molecules containing selenium. Part C. Miscellaneous biological activities. In The Chemistry of Organic Selenium and Tellurium Compounds; Rappoport, Z., Ed.; Wiley: Chichester, UK, 2013; Volume 4, pp. 1119–1174. [Google Scholar]
  10. Santi, C.; Bagnoli, L. Celebrating two centuries of research in selenium chemistry: State of the art and new prospective. Molecules 2017, 22, 2124. [Google Scholar] [CrossRef] [Green Version]
  11. Mugesh, G.; Singh, H.B. Synthetic organoselenium compounds as antioxidants: Glutathione peroxidase activity. Chem. Soc. Rev. 2000, 29, 347–357. [Google Scholar] [CrossRef]
  12. He, X.; Zhong, M.; Li, S.; Li, X.; Li, Y.; Li, Z.; Gao, Y.; Ding, F.; Wen, D.; Lei, Y.; et al. Synthesis and biological evaluation of organoselenium (NSAIDs-SeCN and SeCF3) derivatives as potential anticancer agents. Eur. J. Med. Chem. 2020, 208, 112864. [Google Scholar] [CrossRef] [PubMed]
  13. He, X.; Nie, Y.; Zong, M.; Li, S.; Li, X.; Guo, Y.; Liu, Z.; Gao, Y.; Ding, F.; Wen, D.; et al. New organoselenides (NSAIDs-Se derivatives) as potential anticancer agents: Synthesis, biological evaluation and in silico calculations. Eur. J. Med. Chem. 2021, 218, 113384. [Google Scholar] [CrossRef]
  14. Sabir, S.; Yu, T.T.; Kuppusamy Almohaywi, B.; Iskander, G.; Das, T.; Willcox MD, P.; Black, D.S.; Kumar, N. Novel seleno- and thio-urea containing dihydropyrrol-2-one analogues as antibacterial agents. Antibiotics 2021, 10, 321. [Google Scholar] [CrossRef] [PubMed]
  15. Ali, W.; Benedetti, R.; Handzlik, J.; Zwergel Battisetlli, C. The innovative potential of selenium-containing agents for fighting cancer and viral infections. Drug Discov. Today 2021, 26, 256–263. [Google Scholar] [CrossRef] [PubMed]
  16. Almeida, G.M.; Rafique, J.; Saba, S.; Siminski, T.; Mota NS, R.S.; Filho, D.W.; Braga, A.L.; Pedrosa, R.C.; Ourique, F. Novel selenylated imidazo[1,2-a]pyridines for breast cancer chemotherapy: Inhibition of cell proliferation by Akt-mediated regulation, DNA cleavage and apoptosis. Biochem. Biophys. Res. Commun. 2018, 503, 1291–1297. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, Z.; Lai, H.; Hou, L.; Chen, T. Rational design and action mechanisms of chemically innovative organoselenium in cancer therapy. Chem. Commun. 2020, 56, 179–196. [Google Scholar] [CrossRef]
  18. Santos, D.C.; Rafique, J.; Saba, S.; Almeida, G.M.; Siminski, T.; Padua, C.; Filho, D.W.; Zamoner, A.; Braga, A.L.; Pedrosa, R.C.; et al. Apoptosis oxidative damage-mediated and antiproliferative effect of selenylated imidazo[1,2-a]pyridines on hepatocellular carcinoma HepG2 cells and in vivo. J. Biochem. Mol. Toxicol. 2021, 35, e22663. [Google Scholar] [CrossRef]
  19. Spengler, G.; Gjdacs, M.; Marc, M.A.; Dominguez-Alvarez, E.; Sanmartin, C. Organoselenium compounds as novel adjuvants of chemotherapy drugs—A promising approach to fight cancer drug resistance. Molecules 2019, 24, 336. [Google Scholar] [CrossRef] [Green Version]
  20. Kumawat, A.; Raheem, S.; Ali, F.; Dar, T.A.; Chakrabarty, S.; Rizvi, M.A. Organoselenium compounds as acetylcholinesterase inhibitors: Evidence and mechanism of mixed inhibition. J. Phys. Chem. B 2021, 125, 1531–1541. [Google Scholar] [CrossRef]
  21. Rodrigures, J.; Saba, S.; Joussef, A.C.; Rafique, J.; Braga, A.L. KIO3-catalyzed C(sp2)-H bond selenylation/sulfenylation of (hetero)arenes: Synthesis of chalcogenated (hetero)arenes and their evaluation for anti-Alzheimer activity. Asian J. Org. Chem. 2018, 7, 1819–1824. [Google Scholar] [CrossRef]
  22. Kepp, K.P. Bioinorganic chemistry of Alzheimer’s disease. Chem. Rev. 2012, 112, 5193–5239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Scheide, M.R.; Schneider, A.R.; Jardin GA, M.; Martins, G.M.; Durigon, D.C.; Saba, S.; Rafique, J.; Braga, A.L. Electrochemical synthesis of selenyl-dihydrofurans via anodic selenofunctionalization of allyl-naphthol/phenol derivatives and their anti-Alzheimer activity. Org. Biomol. Chem. 2020, 18, 4916–4921. [Google Scholar] [CrossRef]
  24. Wang, B.; Wang, Z.; Chen, H.; Lu, C.-J.; Li, X. Synthesis and evaluation of 8-hydroxyquinolin derivatives substituted with (benzo[d][1,2]selenazol-3(2H)-one) as effective inhibitor of metal-induced aggregation and antioxidant. Bioorg. Med. Chem. 2016, 24, 4741–4749. [Google Scholar] [CrossRef]
  25. Veloso, I.C.; Delangara, E.; Machado, A.E.; Braga, S.P.; Rosa, G.K.; Bem, A.F.; Rafique, J.; Saba, S.; Trindade, R.N.; Galetto, F.Z.; et al. A selanylimidazopyridine (3-SePh-IP) reverses the prodepressant- and anxiogenic-like effects of a high-fat/high-fructose diet in mice. J. Pharm. Pharmacol. 2021, 73, 673–681. [Google Scholar] [CrossRef]
  26. Gülçin, I.; Trofimov, B.; Kaya, R.; Taslimi, P.; Sobenina, L.; Schmidt, E.; Petrova, O.; Malysheva, S.; Gusarova, N.; Farzaliyev, V.; et al. Synthesis of nitrogen, phosphorus, selenium and sulfur-containing heterocyclic compounds—Determination of their carbonic anhydrase, acetylcholinesterase, butyrylcholinesterase and α-glycosidase inhibition properties. Bioorg. Chem. 2021, 103, 104171. [Google Scholar] [CrossRef]
  27. Martín-Escolano, R.; Etxebeste-Mitxeltorena, M.; Martín-Escolano, J.; Plano, D.; Rosales, M.J.; Espuelas, S.; Moreno, E.; Sánchez-Moreno, M.; Sanmartín, C.; Marín, C. Selenium Derivatives as Promising Therapy for Chagas Disease: In Vitro and In Vivo Studies. ACS Infect. Dis. 2021. [Google Scholar] [CrossRef]
  28. Galant, L.S.; Rafique, J.; Braga, A.L.; Braga, F.C.; Saba, S.; Radi, R.; Rocha JB, T.; Santi CMonsalve, M.; Farina, M.; Bem, A.F. The thiol-modifier effects of organoselenium compounds and their cytoprotective actions in neuronal cells. Neurochem. Res. 2021, 46, 120–130. [Google Scholar] [CrossRef]
  29. Jin, Z.; Du, X.; Xu, Y.; Deng, Y.; Liu, M.; Zhao, Y.; Zhang, B.; Li, X.; Zhang, L.; Peng, C.; et al. Structure of Mpro from COVID-19 virus and discovery of its inhibitors. Nature 2020, 582, 289–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Sies, H.; Parnham, M.J. Potential therapeutic use of ebselen for COVID-19 and other respiratory viral infections. Free Radic. Biol. Med. 2020, 156, 107–112. [Google Scholar] [CrossRef]
  31. Sun, L.-Y.; Chen, C.; Su, J.; Li, J.-Q.; Jiang, Z.; Gao, H.; Ching, J.-Z.; Ding, H.-H.; Zhai, L.; Yang, K.-W. Ebsulfur and Ebselen as highly potent scaffolds for the development of potential SARS-CoV-2 antivirals. Bioorg. Chem. 2021, 112, 104889. [Google Scholar] [CrossRef] [PubMed]
  32. Masuda, R.; Kimura, R.; Karasaki, T.; Sase, S.; Goto, K. Modeling the catalytic cycle of glutathione peroxidase by nuclear magnetic resonance spectroscopic analysis of selenocysteine selenenic acids. J. Am. Chem. Soc. 2021, 143, 6345–6350. [Google Scholar] [CrossRef]
  33. Bortoli, M.; Madabeni, A.; Nogara, P.A.; Omage, F.B.; Ribaudo, G.; Zeppliil, D.; Rocha JB, T.; Orian, L. Chalcogen-nitrogen bond: Insights into a key chemical motif. Catalysts 2020, 11, 114. [Google Scholar] [CrossRef]
  34. Sudati, J.H.; Nogara, P.A.; Saraiva, R.A.; Wagner, C.; Alberto, A.E.; Braga, A.L.; Fachinetto, R.; Piquini, P.C.; Rocha, J.B.T. Diselenoamino acid derivatives as GPx mimics and as substrates of TrxR: In vitro and in silico studies. Org. Biomol. Chem. 2018, 16, 3777–3787. [Google Scholar] [CrossRef] [PubMed]
  35. Rafique, J.; Saba, S.; Canto RF, S.; Frizon TE, A.; Hassan, W.; Waczuk, E.P.; Jan, M.; Back, D.F.; Rocha JB, T.; Braga, A.L. Synthesis and biological evaluation of 2-picolylamide-based diselenides with non-bonded interactions. Molecules 2015, 20, 10095–10109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Collins, C.A.; Fry, F.H.; Holme, A.L.; Yiakouvaki, A.; Al-Qenaei, A.; Pourzand, C.; Jacob, C. Towards multifunctional antioxidants: Synthesis, electrochemistry, in vitro and cell culture evaluation of compounds with ligand/catalytic properties. Org. Biomol. Chem. 2005, 3, 1541–1546. [Google Scholar] [CrossRef]
  37. Nowak, P.; Saluk-Juszczak, J.; Olas, B.; Kolodziejczyk, J.; Wachowics, B. The protective effects of selenoorganic compounds against peroxynitrite-induced changes in plasma proteins and lipids. Cell. Mol. Biol. Lett. 2006, 11, 1–11. [Google Scholar] [CrossRef] [PubMed]
  38. Saluk, J.; Bijak, M.; Nowak, P.; Wachowicz, B. Evaluating the antioxidative activity of diselenide containing compounds in human blood. Bioorg. Chem. 2013, 50, 26–33. [Google Scholar] [CrossRef] [PubMed]
  39. Kloc, K.; Mlochowski, L.; Osajda, K.; Sypeer, L.; Wójtowics, H. New heterocyclic selenenamides: 1,2,4-benzoselenadiazin-3(4H)-ones and 1,3,2-benzodiselenazoles. Tetrahedron Lett. 2002, 43, 4071–4074. [Google Scholar] [CrossRef]
  40. Viglianisi, C.; Simone, L.; Menichetti, S. Copper-Mediated One-Pot Transformation of 2-N-Sulfonyl- aminoaryl Diselenides into Benzo[b][1,4]selenazines. Adv. Synth. Catal. 2012, 354, 77–82. [Google Scholar] [CrossRef]
  41. Radatz, C.S.; Alves, D.; Schneider, P.H. Direct synthesis of 2-aryl-1,3-benzoselenazoles by reaction of bis(2-aminophenyl) diselenides with aryl aldehydes using sodium metabisulfite. Tetrahedron 2013, 69, 1316–1321. [Google Scholar] [CrossRef]
  42. Deobald, A.M.; Camargo LR, S.; Hörner, M.; Rodrigues OE, D.; Alves, D.; Braga, A.L. Synthesis of arylseleno-1,2,3-triazoles via copper-catalyzed 1,3-dipolar cycloaddition of azido arylselenides with alkynes. Synthesis 2011, 15, 2397–2406. [Google Scholar]
  43. Frizon TE, A.; Cararo, J.H.; Saba, S.; Dal-Pont, G.C.; Michels, M.; Braga, H.C.; Pimentel, T.; Dal-Pizzol, F.; Valvassori, S.S.; Rafique, J. Synthesis of novel selenocyanates and evaluation of their effect in cultured mouse neurons submitted to oxidative stress. Oxid. Med. Cell. Longev. 2020, 5417024. [Google Scholar] [CrossRef]
  44. Scheide, M.R.; Peterle, M.M.; Saba, S.; Neto JS, S.; Lenz, G.F.; Cezar, R.D.; Felix, J.F.; Botteselle, G.V.; Schneider, R.; Rafique, J.; et al. Borophosphate glass as an active media for CuO nanoparticle growth: An efficient catalyst for selenylation of oxadiazoles and application in redox reactions. Sci. Rep. 2020, 10, 15233. [Google Scholar] [CrossRef]
  45. Rafique, J.; Farias, G.; Saba, S.; Zapp, E.; Bellettini, I.C.; Salla, C.A.M.; Bechtold, I.H.; Scheide, M.R.; Neto JS, S.; Souza, D.M., Jr.; et al. Selenylated-oxadiazoles as promising DNA intercalators: Synthesis, electronic structure, DNA interaction and cleavage. Dyes Pigm. 2020, 180, 108519. [Google Scholar] [CrossRef]
  46. Peterle, M.M.; Scheide, M.R.; Silva, L.T.; Saba, S.; Rafique, J.; Braga, A.L. Copper-catalyzed three-component reaction of oxadiazoles, elemental Se/S and aryl iodides: Synthesis of chalcogenyl (Se/S)-oxadiazoles. Chemitryselect 2018, 3, 13191–13196. [Google Scholar] [CrossRef]
  47. Silva, L.T.; Azerefo, J.B.; Saba, S.; Rafique, J.; Bortoluzzi, A.J.; Braga, A.L. Solvent- and metal-free chalcogenation of bicyclic arenes Using I2/DMSO as non-metallic catalytic system. Eur. J. Org. Chem. 2017, 32, 4740–4748. [Google Scholar] [CrossRef]
  48. Rafique, J.; Saba, S.; Frizon TE, A.; Braga, A.L. Fe3O4 nanoparticles: A robust and magnetically recoverable catalyst for direct C-H bond selenylation and sulfenylation of benzothiazoles. Chemsitryselect 2018, 3, 328–334. [Google Scholar] [CrossRef]
  49. Rocah MS, T.; Rafique, J.; Saba Azerdo, J.B.; Back, D.; Godoi, M.; Braga, A.L. Regioselective hydrothiolation of terminal acetylene catalyzed by magnetite (Fe3O4) nanoparticles. Synth. Commun. 2017, 47, 291–298. [Google Scholar] [CrossRef]
  50. Rafique, J.; Saba, S.; Rosario, A.R.; Zeni, G.; Braga, A.L. K2CO3-mediated, direct C–H bond selenation and thiolation of 1,3,4-oxadiazoles in the absence of metal catalyst: An eco-friendly approach. RSC Adv. 2014, 4, 51648–51652. [Google Scholar] [CrossRef]
  51. Menichetti, S.; Capperucci Tanini, D.; Braga ALBotteselle, G.V.; Viglianisi, C. A one-pot access to benzo[b][1,4]selenazines from 2-aminoaryl diselenides. Euro. J. Org. Chem. 2016, 18, 3097–3102. [Google Scholar] [CrossRef]
  52. Sharma, U.; Verma, P.K.; Kumar, N.; Kumar, V.; Bala, M.; Singh, B. Phosphane-free green protocol for selective nitro reduction with an iron-based catalyst. Chem. Eur. J. 2011, 17, 5903–5907. [Google Scholar] [CrossRef]
  53. Iwaoka, M.; Tomoda, S. A model study on the effect of an amino group on the antioxidant activity of glutathione peroxidase. J. Am. Chem. Soc. 1994, 116, 2557–2561. [Google Scholar] [CrossRef]
  54. Neese, F. Software update: The ORCA program system, version 4.0. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2017, 8, e1327. [Google Scholar] [CrossRef]
  55. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Glendening, E.D.; Badenhoop, J.K.; Reed, A.E.; Carpenter, J.E.; Bohmann, J.A.; Morales, C.M.; Landis, C.R.; Weinhold, F. NBO 6.0; Theoretical Chemistry Institute, University of Wisconsin: Madison, WI, USA, 2013; Available online: http://nbo6.chem.wisc.edu/ (accessed on 12 July 2021).
Figure 1. Ebselen.
Figure 1. Ebselen.
Molecules 26 04446 g001
Scheme 1. Reaction conditions: (i) Se (3.0 mmol), KOH (6.0 mmol), heated until melted for 5 min and H2O (6.0 mL); (ii) o-halonitrobenzene 1af (1.5 mmol) and THF or DMF (1.5 mL), r.t., 2 h; (iii) bis-nitrobenzene diselenide 2ae (1.5 mmol) and FeSO4.7H2O (5.0 eq), methanol (25.0 mL) and H2O (25.0 mL), reflux, 1 h. NH4OH (15.0 mL), reflux, 10 min.
Scheme 1. Reaction conditions: (i) Se (3.0 mmol), KOH (6.0 mmol), heated until melted for 5 min and H2O (6.0 mL); (ii) o-halonitrobenzene 1af (1.5 mmol) and THF or DMF (1.5 mL), r.t., 2 h; (iii) bis-nitrobenzene diselenide 2ae (1.5 mmol) and FeSO4.7H2O (5.0 eq), methanol (25.0 mL) and H2O (25.0 mL), reflux, 1 h. NH4OH (15.0 mL), reflux, 10 min.
Molecules 26 04446 sch001
Figure 2. (a) UV-Vis spectrum of PhSH oxidation in the presence of H2O2 and diselenide 3b as catalyst. [PhSH] = 10 mmol L−1, [3b] = 0.01 mmol L−1 and [H2O2] = 15 mmol L−1, in methanol at 25 °C; (b) Absorbance plotted against diphenyl disulfide concentration. The red line represents the linear fit. The coefficient of molar absorptivity in 305 nm was 1415 L mol−1 cm−1 (R2 = 0.9996).
Figure 2. (a) UV-Vis spectrum of PhSH oxidation in the presence of H2O2 and diselenide 3b as catalyst. [PhSH] = 10 mmol L−1, [3b] = 0.01 mmol L−1 and [H2O2] = 15 mmol L−1, in methanol at 25 °C; (b) Absorbance plotted against diphenyl disulfide concentration. The red line represents the linear fit. The coefficient of molar absorptivity in 305 nm was 1415 L mol−1 cm−1 (R2 = 0.9996).
Molecules 26 04446 g002
Figure 3. Initial rate (V0) plotted against substrate concentration. The initial rates were calculated from at least two experiments for each concentration of PhSH. The concentrations of 3b and H2O2 were fixed at 1 × 10−5 and 15 × 10−3 mol L−1, respectively. The red line represents the Michaelis–Menten fit.
Figure 3. Initial rate (V0) plotted against substrate concentration. The initial rates were calculated from at least two experiments for each concentration of PhSH. The concentrations of 3b and H2O2 were fixed at 1 × 10−5 and 15 × 10−3 mol L−1, respectively. The red line represents the Michaelis–Menten fit.
Molecules 26 04446 g003
Figure 4. Catalytic efficiency (η) plotted against calculated atomic charges for (a) N atom (qN) of aniline-derived diselenides, (b) Se atom (qSe) of aniline-derived diselenides, and (c) Se atom (qSe) of aniline-derived selenolates.
Figure 4. Catalytic efficiency (η) plotted against calculated atomic charges for (a) N atom (qN) of aniline-derived diselenides, (b) Se atom (qSe) of aniline-derived diselenides, and (c) Se atom (qSe) of aniline-derived selenolates.
Molecules 26 04446 g004
Scheme 2. Reactive intermediates: (a) selenolate (a–e) and (b) selenyl sulfide.
Scheme 2. Reactive intermediates: (a) selenolate (a–e) and (b) selenyl sulfide.
Molecules 26 04446 sch002
Scheme 3. Gram-scale reaction for the synthesis of aniline-derived diselenide 3b.
Scheme 3. Gram-scale reaction for the synthesis of aniline-derived diselenide 3b.
Molecules 26 04446 sch003
Table 1. GPx-like catalytic evaluation of aniline-derived diselenides 3ae.
Table 1. GPx-like catalytic evaluation of aniline-derived diselenides 3ae.
Molecules 26 04446 i001
EntryCatalyst (1 × 10−5 mol L−1)Km (mol L−1)kcat (min−1)η (L mol−1 min−1)
1 Molecules 26 04446 i0020.001700.422248.65
2 Molecules 26 04446 i0030.001140.601527.78
3 Molecules 26 04446 i0040.001340.446333.40
4 Molecules 26 04446 i0050.001051.1851128.57
5 Molecules 26 04446 i0060.001870.470251.51
6 Molecules 26 04446 i0070.000810.405500.11
7 Molecules 26 04446 i0080.000880.9181044.19
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Botteselle, G.V.; Elias, W.C.; Bettanin, L.; Canto, R.F.S.; Salin, D.N.O.; Barbosa, F.A.R.; Saba, S.; Gallardo, H.; Ciancaleoni, G.; Domingos, J.B.; et al. Catalytic Antioxidant Activity of Bis-Aniline-Derived Diselenides as GPx Mimics. Molecules 2021, 26, 4446. https://doi.org/10.3390/molecules26154446

AMA Style

Botteselle GV, Elias WC, Bettanin L, Canto RFS, Salin DNO, Barbosa FAR, Saba S, Gallardo H, Ciancaleoni G, Domingos JB, et al. Catalytic Antioxidant Activity of Bis-Aniline-Derived Diselenides as GPx Mimics. Molecules. 2021; 26(15):4446. https://doi.org/10.3390/molecules26154446

Chicago/Turabian Style

Botteselle, Giancarlo V., Welman C. Elias, Luana Bettanin, Rômulo F. S. Canto, Drielly N. O. Salin, Flavio A. R. Barbosa, Sumbal Saba, Hugo Gallardo, Gianluca Ciancaleoni, Josiel B. Domingos, and et al. 2021. "Catalytic Antioxidant Activity of Bis-Aniline-Derived Diselenides as GPx Mimics" Molecules 26, no. 15: 4446. https://doi.org/10.3390/molecules26154446

Article Metrics

Back to TopTop