Next Article in Journal
Bis(2,2,6,6-tetramethyl-1-(λ1-oxidaneyl)piperidin-4-yl) 3,3′-((2-hydroxyethyl)azanediyl)dipropionate
Previous Article in Journal
Bis(N-tert-butylacetamido)(dimethylamido)(chloro)titanium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Synthesis and Monoamine Oxidase Inhibition Properties of 4-(2-Methyloxazol-4-yl)benzenesulfonamide

1
Pharmaceutical Technology Transfer Center, Yaroslavl State Pedagogical University Named after K. D. Ushinsky, Yaroslavl 150000, Russia
2
Department of Organic Chemistry, Russian State University Named after A. N. Kosygin, Moscow 115035, Russia
3
Pharmaceutical Chemistry and Centre of Excellence for Pharmaceutical Sciences, North-West University, Potchefstroom 2520, South Africa
*
Author to whom correspondence should be addressed.
Molbank 2024, 2024(1), M1787; https://doi.org/10.3390/M1787
Submission received: 9 January 2024 / Revised: 19 February 2024 / Accepted: 23 February 2024 / Published: 6 March 2024

Abstract

:
4-(2-Methyloxazol-4-yl)benzenesulfonamide was synthesized by the reaction of 4-(2-bromoacetyl)benzenesulfonamide with an excess of acetamide. The compound was evaluated as a potential inhibitor of human monoamine oxidase (MAO) A and B and was found to inhibit these enzymes with IC50 values of 43.3 and 3.47 μM, respectively. The potential binding orientation and interactions of the inhibitor with MAO-B were examined by molecular docking, and it was found that the sulfonamide group binds and interacts with residues of the substrate cavity. 4-(2-Methyloxazol-4-yl)benzenesulfonamide showed no cytotoxic effect against human stromal bone cell line (HS-5) in the concentration range of 1–100 µmol. Thus, the new selective MAO-B inhibitor was identified, which may be used as the lead compound for the development of antiparkinsonian agents.

1. Introduction

Neurodegenerative disorders significantly affect the health of the human population and place a burden on economies of countries around the world. Among these disorders, Parkinson’s disease (PD) is the second most prevalent and is associated with the death of dopaminergic motor neurons in the brain. At present, PD is not curable; however, the motor symptoms associated with this disorder are effectively treated with levodopa, the metabolic precursor of dopamine. To enhance the therapeutic action of levodopa, this drug is frequently combined with monoamine oxidase (MAO) B inhibitors, compounds that reduce the MAO-mediated central metabolism of dopamine [1]. While MAO-B inhibitors may allow for a reduction in the effective levodopa dosage, these compounds have also been studied as neuroprotective agents [2], candidates for reducing neuroinflammation [3,4,5,6], as well as compounds with potential value in the therapy of oncological diseases [7]. Possible mechanisms by which MAO-B inhibitors may exert neuroprotective effects include the enhancement of the levels of brain-derived neurotrophic factors (BDNFs) [8] and the molecular adhesion of L1CAM (L1) nerve cells, which can also promote axonal regrowth and increase neuronal survival, synaptic plasticity, and remyelination [9]. Thus, the development of a new generation of neuroprotective agents that act by inhibiting MAO might have relevance in the future treatment of neurodegenerative disorders.
Benzenesulfonamide compounds have been identified as potent and isoform-specific inhibitors of MAO-B, with some compounds exhibiting potencies in the nanomolar range. Such compounds might represent promising candidates for the future treatment of PD [10,11]. Also, it has been established that 1,3-oxazole derivatives exhibit the potent and specific inhibition of the MAO-B isoform [12]. This work reports a successful attempt to combine both these structural features into a single molecule (Figure 1).
Recently, our research group, as well as others, reported a variety of new lead compounds for the development of isoform-specific MAO-B inhibitors, such as 2,1-benzisoxazoles [13,14], indazoles [15,16], and quinoxalines [17,18]. Continuing the search for novel MAO inhibitors, in this work, we used a simple convergent approach to synthesize a new 1,3-oxazole compound substituted with a primary benzenesulfonamide functionality.

2. Results

2.1. Chemistry

We investigated the possibility of synthesizing the target oxazole 1 by the reaction of sulfonamide-containing phenacyl bromide 2 with acetamide or ammonium acetate.
The starting material, phenacyl bromide 3, was obtained according to a well-known three-step procedure using 4-acetylbenzenesulfonamide (2) as a key reagent (Scheme 1) [19,20]. The best yield (63%) of the target oxazole 1 was obtained by fusing phenacyl bromide 3 with an excess of acetamide at 150 °C ((Scheme 2). It was shown that the reaction ends in 15–20 min, and longer heating or increasing the reaction temperature to 190 °C leads to a sharp decrease in the yield of 1 and the appearance of degradation by-products.
Additionally, when we used the method in [21] based on reaction compound 5 with ammonium acetate in acetic acid, the target oxazole 1 was obtained at a much lower yield (24%).

2.2. MAO Inhibition

The MAO inhibition potency of 4-(2-methyloxazol-4-yl)benzenesulfonamide (1) was evaluated using recombinant human MAO-A and MAO-B following the protocol described in the literature [13]. The results of the MAO inhibition studies are presented in Table 1. Compound 1 inhibited MAO-B with an IC50 value of 3.47 μM, whereas weak inhibition of the MAO-A isoform was recorded.
The potential orientations by which 1 binds to the active site of human MAO-B was investigated by molecular docking using the CDOCKER module of the Discovery Studio 3.1 modeling software. Docking predicts that the inhibitor binds with the sulfonamide group placed in the substrate cavity of the enzyme, while the oxazole moiety extends towards the entrance cavity of the MAO-B active site (Figure 2).
The experimentally determined binding mode of zonisamide is also shown [22]. The sulfonamide group is placed in the same space as that of zonisamide, as determined by X-ray crystallography [22]. Hydrogen bonding occurs between the sulfonamide functional group and an active site water and Gln-206. Pi–sulfur interactions form between Tyr-60 and the sulfonamide functional group, as well between Cys-172 and the oxazole moiety. A pi-pi interaction also occurs between the oxazole moiety and Tyr-326. It is interesting to note that the oxazole moiety does not protrude deep into the entrance cavity and therefore cannot be viewed as a cavity-spanning MAO-B inhibitor. Compounds that bind to both the substrate and entrance cavities often exhibit submicromolar potencies due to the additional stabilization afforded by nonpolar interactions with the lipophilic environment of the entrance cavity [23]. This may explain the moderate MAO-B inhibition potency observed for compound 1.
Table 1. The inhibition of human MAO-A and MAO-B by 4-(2-methyloxazol-4-yl)benzenesulfonamide (1).
Table 1. The inhibition of human MAO-A and MAO-B by 4-(2-methyloxazol-4-yl)benzenesulfonamide (1).
StructureIC50 (μM ± SD) 1
MAO-AMAO-B
Molbank 2024 m1787 i00143.3 ± 7.123.47 ± 0.31
Curcumin [24] 2 5.02 ± 0.452.56 ± 0.21
Toloxatone [25] 23.92 -
1 The IC50 values are presented as the means ± standard deviation (SD) of triplicate measurements. 2 Reference inhibitors.

2.3. Cytotoxicity In Vitro

The cytotoxicity of compound 1 was investigated on human cell line HS-5 (human bone marrow stroma). In this respect, the cell viability after being incubated with 1–100 µM of the test compound was evaluated using the MTT assay [26,27].
The result of that experiment showed that compound 1 did not exhibit toxicity at the concentration range used for this study (Figure 3).

3. Discussion

This study reports the MAO inhibition potency of 4-(2-methyloxazol-4-yl)benzenesulfonamide (1). This compound inhibited MAO-B with an IC50 value of 3.47 μM, while the MAO-A isoform was inhibited with an IC50 value of 43.3 μM. Molecular docking experiments show that the inhibitor is restricted to the substrate cavity and does not extend deep into the entrance cavity, which may explain the moderate MAO-B inhibition potency observed for this compound. The discovery of this active MAO-B inhibitor paves the way for the future discovery of MAO-B inhibitors among 1,3-oxazole derivatives substituted with a primary benzenesulfonamide. Such compounds may find application in the treatment of neurodegenerative disorders such as PD.

4. Materials and Methods

4.1. General

All reagents and solvents were obtained from commercial sources (Aldrich, Merck, Aladdin, Gernsheim, Germany) and were used without purification. Reactions were monitored by analytical thin-layer chromatography (TLC) using Merck TLC sheets. Visualization of the developed sheets was performed by fluorescence quenching at 254 nm. 1H NMR and 13C NMR spectra were recorded on a Varian 400 Unity Plus instrument (400 MHz for 1H and 100 MHz for 13C, respectively). Chemical shifts (δ) are given in parts per million (ppm) and were referenced to the solvent signal for DMSO-d6 (2.50 ppm for proton and 39.52 ppm for carbon), while the coupling constants (J) were reported in hertz (Hz). Multiplicities are abbreviated as follows: s = singlet, d = doublet, dd = doublet of doublets, t = triplet, q = quartet, m = multiplet. Melting points were determined on an Electrothermal IA 9300 series digital melting point apparatus. Mass spectra were recorded using ESI ionization on a microTOF spectrometer (Bruker Daltonics Inc., Bremen, Germany).

4.2. Procedure for the Preparation of 4-(2-Bromoacetyl)benzenesulfonamide (3)

Compound 3 was prepared according the method described in [20] using 0.90 g 4-acetylbenzenesulfonamide (2) as a starting material. (3) In total, 1.22 g (98%) of 4-(2-Bromoacetyl)benzenesulfonamide was isolated as the beige crystalline solid. mp 154–155 °C. 1H NMR (400 MHz, DMSO-d6) δ 8.17 (d, J = 8.2 Hz, 2H), 7.96 (d, J = 8.2 Hz, 2H), 7.57 (s, 2H), 4.98 (s, 2H).

4.3. Synthesis and Characterization of 4-(2-Methyloxazol-4-yl)benzenesulfonamide (1)

Acetamide (0.12 g, 2 mmol, 3 equiv.) and 4-(2-bromoacetyl)benzenesulfonamide (3; 0.19 g, 0.68 mmol, 1 equiv.) were mixed together. The reaction mixture was melted, stirred at 150 °C for 20 min, and subsequently diluted with cold water (30 mL). The resulting precipitate was collected by filtration, washed with water (10 mL), and air-dried at 50 °C. Yield 0.101 g, 63%, beige solid, mp 237–239 °C; 1H NMR (400 MHz, DMSO-d6) δ 8.61 (s, 1H), 7.93 (d, J = 8.4 Hz, 2H), 7.85 (d, J = 8.4 Hz, 2H), 7.38 (s, 2H), 2.48 (s, 3H); 13C NMR (101 MHz, DMSO-d6) δ 162.6, 143.7, 139.2, 137.0, 134.9, 126.9, 125.9, 14.2; MS (ESI+): m/z [M+H]+ Anal. Calcd for C10H11N2O3S: 239.0485. Found: 239.0489 (Supplementary Materials).

4.4. MAO Inhibition Studies

The measurement of IC50 values for the inhibition of human MAO-A and MAO-B was carried out according to a previously reported protocol [13]. Recombinant human MAO-A and MAO-B were obtained from Sigma-Aldrich (USA), and fluorescence measurements were recorded with a SpectraMax iD3 multi-mode microplate reader (Molecular Devices). The measurement of MAO activity was based on the fluorescence signal generated when the substrate, kynuramine, was oxidized by the MAOs to yield 4-hydroxyquinoline.

4.5. Molecular Docking Studies

Molecular docking was carried out according to the previously reported protocol using the Discovery Studio 3.1 suite of software [11]. The X-ray crystal structure of MAO-B (PDB code: 2V5Z) complexed to safinamide was used for the docking study [28]. The illustration was created with the PyMOL molecular graphics system [29].

4.6. Cytotoxicity Assay

The human cell culture line HS-5 (bone marrow stroma) was grown in a mixture (1:1) of RPMI-1640 and Ham’s F12 media without glutamine. To the media, we added FBS (10%), L-glutamine (2 mM), penicillin (50 IU/mL), and streptomycin (50 μg/mL). The test compounds were dissolved in DMSO and added to the culture media to yield a final concentration of 0.1% DMSO. As a positive control (0% cell viability), the cells were exposed to sodium azide (0.1%). As a negative control, cell viability was measured in the absence of the test compound. The cytotoxicity of the test compound was determined by the MTT [3-(4,5-dimethylthiazole-2-yl)-2,5-diphenyl tetrazolium bromide] protocol [27]. Absorbance was measured at a wavelength of 590 nm with a CLARIOstar (BMG Labtech) spectrophotometer. The blank consisted of culture medium.

Supplementary Materials

1H- and 13C-NMR spectra copies of synthesized compounds can be found in the File S1 (Supplementary Materials).

Author Contributions

Conceptualization of the study was conducted by A.A.S.; formal analysis and investigation by J.A.E. and A.P.; writing—original draft preparation was conducted by A.A.S.; writing—review and editing was conducted by M.K.K. and J.P.P. All authors have read and agreed to the published version of the manuscript.

Funding

The chemical section of this work was supported by the Russian Science Foundation (project 22-13-20085). The MAO inhibition studies were funded by the National Research Foundation of South Africa [grant specific unique reference number (UID) 137997 (JPP)]. The grant holders acknowledge that the opinions, findings, and conclusions or recommendations expressed in any publication generated by NRF-supported research are that of the authors and that the NRF accepts no liability whatsoever in this regard.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no competing interests.

References

  1. Tan, Y.Y.; Jenner, P.; Chen, S.D. Monoamine Oxidase-B Inhibitors for the Treatment of Parkinson’s Disease: Past, Present, and Future. J. Park. Dis. 2022, 12, 477–493. [Google Scholar] [CrossRef]
  2. Szökő, É.; Tábi, T.; Riederer, P.; Vécsei, L.; Magyar, K. Pharmacological aspects of the neuroprotective effects of irreversible MAO-B inhibitors, selegiline and rasagiline, in Parkinson’s disease. J. Neural Transm. 2018, 125, 1735–1749. [Google Scholar] [CrossRef]
  3. Ostadkarampour, M.; Putnins, E.E. Monoamine Oxidase Inhibitors: A Review of Their Anti-Inflammatory Therapeutic Potential and Mechanisms of Action. Front. Pharmacol. 2021, 12, 676239. [Google Scholar] [CrossRef]
  4. Santos, M.A.; Chand, K.; Chaves, S. Recent progress in multifunctional metal chelators as potential drugs for Alzheimer’s disease. Coord. Chem. Rev. 2016, 327, 287–303. [Google Scholar] [CrossRef]
  5. Meredith, G.E.; Totterdell, S.; Beales, M.; Meshul, C.K. Impaired glutamate homeostasis and programmed cell death in a chronic MPTP mouse model of Parkinson’s disease. Exp. Neurol. 2009, 219, 334–340. [Google Scholar] [CrossRef]
  6. Belaidi, A.A.; Bush, A.I. Iron neurochemistry in Alzheimer’s disease and Parkinson’s disease: Targets for therapeutics. J. Neurochem. 2016, 139, 179–197. [Google Scholar] [CrossRef]
  7. Zarmouh, N.O.; Messeha, S.S.; Mateeva, N.; Gangapuram, M.; Flowers, K.; Eyunni, S.V.K.; Zhang, W.; Redda, K.K.; Soliman, K.F.A. The Antiproliferative Effects of Flavonoid MAO Inhibitors on Prostate Cancer Cells. Molecules 2020, 25, 2257. [Google Scholar] [CrossRef]
  8. Balu, D.T.; Hoshaw, B.A.; Malberg, J.E.; Rosenzweig-Lipson, S.; Schechter, L.E.; Lucki, I. Differential regulation of central BDNF protein levels by antidepressant and non-antidepressant drug treatments. Brain Res. 2008, 1211, 37–43. [Google Scholar] [CrossRef] [PubMed]
  9. Li, R.; Sahu, S.; Schachner, M. Phenelzine, a cell adhesion molecule L1 mimetic small organic compound, promotes functional recovery and axonal regrowth in spinal cord-injured zebrafish. Pharmacol. Biochem. Behav. 2018, 171, 30–38. [Google Scholar] [CrossRef] [PubMed]
  10. Grover, N.D.; Limaye, R.P.; Gokhale, D.V.; Patil, T.R. Zonisamide: A review of the clinical and experimental evidence for its use in Parkinson’s disease”. Ind. J. Pharmacol. 2013, 45, 547–555. [Google Scholar] [CrossRef] [PubMed]
  11. Shetnev, A.; Shlenev, R.; Efimova, J.; Ivanovskii, S.; Tarasov, A.; Petzer, A.; Petzer, J.P. 1,3,4-Oxadiazol-2-ylbenzenesulfonamides as privileged structures for the inhibition of monoamine oxidase B. Bioorg. Med. Chem. Lett. 2019, 29, 126677. [Google Scholar] [CrossRef] [PubMed]
  12. Qazi, S.U.; Naz, A.; Hameed, A.; Osra, F.A.; Jalil, S.; Iqbal, J.; Shah, S.A.; Mirza, A.Z. Semicarbazones, thiosemicarbazone, thiazole and oxazole analogues as monoamine oxidase inhibitors: Synthesis, characterization, biological evaluation, molecular docking, and kinetic studies. Bioorg. Chem. 2021, 115, 105209. [Google Scholar] [CrossRef] [PubMed]
  13. Shetnev, A.; Kotov, A.; Kunichkina, A.; Proskurina, I.; Baykov, S.; Korsakov, M.; Petzer, A.; Petzer, J.P. Monoamine oxidase inhibition properties of 2,1-benzisoxazole derivatives. Mol. Divers. 2023. [Google Scholar] [CrossRef] [PubMed]
  14. Agrawal, N.; Mishra, P. Synthesis, monoamine oxidase inhibitory activity and computational study of novel isoxazole derivatives as potential antiparkinson agents. Comput. Biol. Chem. 2019, 79, 63–72. [Google Scholar] [CrossRef] [PubMed]
  15. Tzvetkov, N.T.; Antonov, L. Subnanomolar indazole-5-carboxamide inhibitors of monoamine oxidase B (MAO-B) continued: Indications of iron binding, experimental evidence for optimised solubility and brain penetration. J. Enzym. Inhib. Med. Chem. 2017, 32, 960–967. [Google Scholar] [CrossRef] [PubMed]
  16. Efimova, J.A.; Shetnev, A.A.; Baykov, S.V.; Petzer, A.; Petzer, J.P. 3-(3,4-Dichlorophenyl)-5-(1H-indol-5-yl)-1,2,4-oxadiazole. Molbank 2023, 2023, M1552. [Google Scholar] [CrossRef]
  17. Panova, V.A.; Filimonov, S.I.; Chirkova, Z.V.; Kabanova, M.V.; Shetnev, A.A.; Korsakov, M.K.; Petzer, A.; Petzer, J.P.; Suponitsky, K.Y. Investigation of pyrazolo[1,5-a]quinoxalin-4-ones as novel monoamine oxidase inhibitors. Bioorg. Chem. 2021, 108, 104563. [Google Scholar] [CrossRef]
  18. Khattab, S.N.; Hassan, S.Y.; Bekhit, A.A.; El Massry, A.M.; Langer, V.; Amer, A. Synthesis of new series of quinoxaline-based MAO-inhibitors and docking studies. Eur. J. Med. Chem. 2010, 45, 4479–4489. [Google Scholar] [CrossRef]
  19. Jacobs, J.W.; Leadbetter, M.R.; Bell, N.; Koo-McCoy, S.; Carreras, C.W.; He, L.; Kohler, J.; Kozuka, K.; Labonté, E.D.; Navre, M.; et al. Discovery of tenapanor: A first-in-class minimally systemic inhibitor of intestinal Na+/H+ exchanger isoform 3. ACS Med. Chem. Lett. 2022, 13, 1043–1051. [Google Scholar] [CrossRef]
  20. Scobie, M.; Wallner, O.; Koolmeister, T.; Vallin, K.S.A.; Henriksson, C.M.; Homan, E.; Helleday, T.; Jacques, S.; Desroses, M.; Jacques-Cordonnier, M.-C.; et al. MTH1 Inhibitors for Treatment of Inflammatory and Autoimmune Conditions. U.S. Patent 16,102,170, 31 August 2018. [Google Scholar]
  21. Denney, D.B.; Pastor, S.D. Structures in solution of adducts of hexamethylphosphorus triamide and substituded benzils. Phosphorus Sulfur. Silicon Relat. Elem. 1983, 16, 239–246. [Google Scholar] [CrossRef]
  22. Binda, C.; Aldeco, M.; Mattevi, A.; Edmondson, D.E. Interactions of monoamine oxidases with the antiepileptic drug zo-nisamide: Specificity of inhibition and structure of the human monoamine oxidase B complex. J. Med. Chem. 2011, 54, 909–912. [Google Scholar] [CrossRef]
  23. Hubalek, F.; Binda, C.; Khalil, A.; Li, M.; Mattevi, A.; Castagnoli, N.; Edmondson, D.E. Demonstration of isoleucine 199 as a structural determinant for the selective inhibition of human monoamine oxidase B by specific reversible inhibitors. J. Biol. Chem. 2005, 280, 15761–15766. [Google Scholar] [CrossRef] [PubMed]
  24. Khatri, D.K.; Juvekar, A.R. Kinetics of Inhibition of Monoamine Oxidase Using Curcumin and Ellagic Acid. Pharmacogn. Mag. 2016, 12, 116–120. [Google Scholar] [CrossRef]
  25. Berlin, I.; Zimmer, R.; Thiede, H.; Payan, C.; Hergueta, T.; Robin, L.; Puech, A. Comparison of the monoamine oxidase inhibiting properties of two reversible and selective monoamine oxidase-A inhibitors moclobemide and toloxatone, and assessment of their effect on psychometric performance in healthy subjects. Br. J. Clin. Pharmacol. 1990, 30, 805–816. [Google Scholar] [CrossRef] [PubMed]
  26. Ghasemi, M.; Turnbull, T.; Sebastian, S.; Kempson, I. The MTT Assay: Utility, Limitations, Pitfalls, and Interpretation in Bulk and Single-Cell Analysis. Int. J. Mol. Sci. 2021, 22, 12827. [Google Scholar] [CrossRef]
  27. Mosmann, T. Rapid colorimetric assay for cellular growth and survival: Application to proliferation and cytotoxicity assays. J. Immunol. Methods 1983, 65, 55–63. [Google Scholar] [CrossRef]
  28. Binda, C.; Wang, J.; Pisani, L.; Caccia, C.; Carotti, A.; Salvati, P.; Edmondson, D.E.; Mattevi, A. Structures of human monoamine oxidase B complexes with selective noncovalent inhibitors: Safinamide and coumarin analogs. J. Med. Chem. 2007, 50, 5848–5852. [Google Scholar] [CrossRef]
  29. DeLano, W.L. The PyMOL Molecular Graphics System; DeLano Scientific: San Carlos, CA, USA, 2002. [Google Scholar]
Figure 1. Structures of MAO inhibitors containing the sulfonamide or 1,3-oxazole moieties [11,12].
Figure 1. Structures of MAO inhibitors containing the sulfonamide or 1,3-oxazole moieties [11,12].
Molbank 2024 m1787 g001
Scheme 1. Synthesis of 4-(2-bromoacetyl)benzenesulfonamide (3).
Scheme 1. Synthesis of 4-(2-bromoacetyl)benzenesulfonamide (3).
Molbank 2024 m1787 sch001
Scheme 2. Synthesis of 4-(2-methyloxazol-4-yl)benzenesulfonamide (1).
Scheme 2. Synthesis of 4-(2-methyloxazol-4-yl)benzenesulfonamide (1).
Molbank 2024 m1787 sch002
Figure 2. The predicted binding orientation of compound 1 to the active site of MAO-B. Zonisamide shown in pink, oxazole compound 1 shown in blue, red dots showing the binding regions to the target cavity.
Figure 2. The predicted binding orientation of compound 1 to the active site of MAO-B. Zonisamide shown in pink, oxazole compound 1 shown in blue, red dots showing the binding regions to the target cavity.
Molbank 2024 m1787 g002
Figure 3. Cell viability after treatment of HS-5 stromal cell line with compound 1 (1–100 µM).
Figure 3. Cell viability after treatment of HS-5 stromal cell line with compound 1 (1–100 µM).
Molbank 2024 m1787 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shetnev, A.A.; Efimova, J.A.; Korsakov, M.K.; Petzer, A.; Petzer, J.P. Synthesis and Monoamine Oxidase Inhibition Properties of 4-(2-Methyloxazol-4-yl)benzenesulfonamide. Molbank 2024, 2024, M1787. https://doi.org/10.3390/M1787

AMA Style

Shetnev AA, Efimova JA, Korsakov MK, Petzer A, Petzer JP. Synthesis and Monoamine Oxidase Inhibition Properties of 4-(2-Methyloxazol-4-yl)benzenesulfonamide. Molbank. 2024; 2024(1):M1787. https://doi.org/10.3390/M1787

Chicago/Turabian Style

Shetnev, Anton A., Julia A. Efimova, Mikhail K. Korsakov, Anél Petzer, and Jacobus P. Petzer. 2024. "Synthesis and Monoamine Oxidase Inhibition Properties of 4-(2-Methyloxazol-4-yl)benzenesulfonamide" Molbank 2024, no. 1: M1787. https://doi.org/10.3390/M1787

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