Next Article in Journal
2,2’-[(1E,1’E)-{[Hexa-2,4-diyne-1,6-diylbis(oxy)]bis(2,1-phenylene)}bis(ethene-2,1-diyl)]bis(4H-chromen-4-one)
Previous Article in Journal
2-(2,7-Bis(pyridin-3-ylethynyl)fluoren-9-ylidene)malononitrile
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Short Note

3-(4-Fluorophenyl)-1-(1-(4-fluorophenyl)-3,3,3-trifluoroprop-1-en-1-yl)-5-fluoro-1H-pyrazole

by
Stanislav A. Paveliev
,
Alexander O. Ustyuzhanin
,
Igor B. Krylov
* and
Alexander O. Terent’ev
*
N. D. Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences, 47 Leninsky Prospekt, 119991 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Molbank 2023, 2023(2), M1620; https://doi.org/10.3390/M1620
Submission received: 6 March 2023 / Revised: 5 April 2023 / Accepted: 11 April 2023 / Published: 13 April 2023
(This article belongs to the Section Organic Synthesis)

Abstract

:
In this work, the title compound was synthesized via the visible-light-induced radical denitrogenative trifluoromethylation of the corresponding vinyl azide followed by Cs2CO3-mediated defluorinative cyclization of the resultant azine. The widely available sodium trifluoromethanesulfinate is used as a precursor of CF3 radicals, while graphitic carbon nitride (g-C3N4) is employed as an environmentally friendly, cheap, and efficient heterogeneous photocatalyst. The structure of the synthesized compound was established by 1H, 13C, 19F-NMR, IR spectroscopy, and mass-spectrometry.

1. Introduction

The development of new methods for the synthesis of organofluorine compounds is among the most urgent tasks of modern organic and medicinal chemistry [1,2]. Since the middle of the last century, organofluorine bioactive compounds have accounted for 5–15% of the annual number of new drugs approved on the world pharmaceutical market [3,4]. Fluorinated heterocyclic compounds [5,6], including pyrazoles [7], are often found among biorelevant structures. In this regard, the demand for the incorporation of fluorine atoms and fluorine-containing functional groups into organic molecules is constantly growing.
One of the most powerful and widely used tools in the chemistry of organofluorine compounds is radical fluoroalkylation [8,9] and, in particular, trifluoromethylation. It is known that the introduction of the CF3 group into a molecule can enhance its ability to permeate through biological membranes, increase the strength of binding to target receptors, and improve metabolic stability [10]. Recently, significant advances were achieved in the field of photochemical radical fluoroalkylation [11,12]. Photoredox catalysis allows one to selectively introduce fluoroalkyl moiety under mild conditions with the use of available and stable precursors of fluorinated radicals [13]. Therefore, the search for new efficient photocatalytic systems is quite topical.
As an alternative to traditional homogeneous (molecular) photocatalysis, heterogeneous photocatalysis on semiconductors is gaining increasing attention [14]. Visible light-induced heterogeneous photocatalysis with semiconducting graphitic carbon nitride (g-C3N4) provided an excellent platform for the introduction of the CF3 group into various classes of substrates, including heterocyclic compounds [15,16]. g-C3N4 is characterized by its stability, ease of production and separation, high efficiency under visible light irradiation, and powerful oxidizing and reduction abilities [17,18].
In this work, a one-pot synthetic radical transformation of vinyl azide 1 leading to the formation of fluorinated trisubstituted pyrazole 2 was implemented (Scheme 1).

2. Results and Discussion

At the first stage, the interaction between 1-(1-azidovinyl)-4-fluorobenzene 1 and sodium trifluoromethanesulfinate (Langlois reagent) under the g-C3N4-mediated heterogeneous photocatalytic conditions was performed (Scheme 2). The reaction proceeds under irradiation with a high-power blue (450 nm) LED lamp in an air atmosphere in DMSO. In this process, the photoexcited g-C3N4 acts as an electron acceptor due to its hole-electron semiconductor nature. Oxidation of the CF3SO2 anion leads to the formation of trifluoromethyl radicals with the extrusion of the SO2 molecule. Then, the CF3 radical is added to the terminal carbon atom of the C=C double bond of vinyl azide 1 [19]. This addition leads to the elimination of nitrogen molecule and the formation of N-centered iminyl radicals prone to dimerization with the formation of an N–N bond [20].
The resulting intermediate azine was used in the next step without isolation and purification. The cyclization reaction was carried out for 2 h at 100 °C in DMSO. Cesium carbonate was used as the base. Transformation proceeds with the elimination of two hydrogen fluoride molecules from the azine molecule [21]. The structure of compound 2 was confirmed by 1H, 13C, and 19F NMR spectroscopy, high-resolution mass spectrometry, and FT-IR spectroscopy (Figures S1–S5, Supplementary Materials).

3. Materials and Methods

3.1. General

1-(1-azidovinyl)-4-fluorobenzene (1) [22] and g-C3N4 [23] were prepared according to the published methods. All commercially available reagents and solvents were used without purification. 1H, 13C NMR, and 19F NMR spectra were taken with a Bruker AM-300 machine (Bruker AXS Handheld Inc., Kennewick, WA, USA) (at frequencies of 300, 75, and 282 MHz) in CDCl3 using the residual solvent peak as a reference. J values are given in Hz. The high-resolution mass spectrum was measured on a Bruker microTOF II instrument using electrospray ionization (ESI). FT-IR spectra were recorded on a Bruker ALPHA FT-IR spectrometer. The TLC analysis was carried out on silica gel chromatography plates Macherey-Nagel Alugram UV254 (Macherey-Nagel GmbH & Co. KG, Düren, Germany). The melting points were determined on a Stuart SMP10 Kofler hot-stage apparatus (Stuart Scientific Co. Ltd., Staffordshire, UK). Chromatography of final product was performed on silica gel (0.060–0.200 mm, 60 Å, CAS 7631-86-9).

3.2. 3-(4-Fluorophenyl)-1-(1-(4-Fluorophenyl)-3,3,3-Trifluoroprop-1-en-1-yl)-5-Fluoro-1H-Pyrazole (2)

A 10 mL test tube was charged with vinyl azide (1) (82 mg, 0.5 mmol), sodium trifluoromethanesulfinate (156 mg, 1.0 mmol), g-C3N4 (30 mg), and DMSO (5 mL). Reaction mixture was stirred for 24 h at 28−30 °C under air atmosphere while irradiated with 36 W 450 nm LED. After that, Cs2SO3 (0.6 mmol, 196 mg) was added, and reaction mixture was stirred at 100 °C for 2 h. Upon completion, the reaction mixture was cooled to room temperature, diluted with water (100 mL), and extracted with DCM (3 × 20 mL). Combined organic phases were dried over anhydrous Na2SO4, and rotary evaporated. The crude product was purified by column chromatography (hexane/dichloromethane, 1:1, v/v) to afford 77 mg (42%) of target compound (2) as a light-yellow oil, Rf = 0.50 (hexane/dichloromethane, 1:1, v/v). 1H NMR (CDCl3, ppm): δ 7.75 (dd, J = 8.8, 5.4, 2H), 7.43 (dd, J = 8.8, 5.4, 2H), 7.20–7.05 (m, 4H), 6.46 (q, J = 7.9, 1H), and 6.13 (d, J = 5.1, 1H). 13C NMR (CDCl3, ppm): δ 164.1 (d, J = 251.2), 163.6 (d, J = 248.8), 158.7 (d, J = 282.9), 151.1 (d, J = 10.4), 131.3 (dq, J = 9.1, 1.5), 129.0, 127.6 (d, J = 8.4), 127.1, 123.0 (q, J = 271.3), 116.0 (d, J = 21.8), 115.7 (d, J = 22.1), 110.2 (q, J = 35.7), and 87.0 (d, J = 14.2). 19F NMR (CDCl3, ppm): δ −54.62 (d, J = 8.4), −109.34 (tt, J = 8.4, 5.1), −111.89 (tt, J = 8.4, 5.1), and −122.44 (d, J = 8.4). IR spectrum, ν, cm−1: 2104, 1662, 1607, 1586, 1530, 1512, 1477, 1447, 1364, 1265, 1234, 1160, 1123, and 841. HRMS (ESI-TOF), m/z: calcd for C18H10F6N2 [M + H]+, 369.0821, found, 369.0821.

4. Conclusions

Previously unknown 3-(4-fluorophenyl)-1-(1-(4-fluorophenyl)-3,3,3-trifluoroprop-1-en-1-yl)-5-fluoro-1H-pyrazole (2) was obtained with a yield of 42% via a one-pot synthetic sequence, including g-C3N4-catalyzed heterogeneous photochemical trifluoromethylation of vinyl azide 1 with sodium trifluoromethanesulfinate as the CF3-radical source under visible-light irradiation and defluorinative cyclization of intermediate azine under basic conditions. The structure of the compound was confirmed by NMR spectroscopy, mass spectrometry, and IR analysis.

Supplementary Materials

Figure S1: 1H NMR spectra of 2, Figure S2: 13C NMR spectra of 2, Figure S3: 19F NMR spectra of 2, Figure S4: HRMS (ESI) of 2, Figure S5: FT-IR spectra of 2.

Author Contributions

Conceptualization, A.O.T. and I.B.K.; methodology, S.A.P. and A.O.U.; writing—original draft preparation, A.O.U.; writing—review and editing, S.A.P.; supervision, A.O.T. All authors have read and agreed to the published version of the manuscript.

Funding

This study was performed with financial support from the Russian Science Foundation (grant No. 22-73-00083).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, J.; Sánchez-Roselló, M.; Aceña, J.L.; del Pozo, C.; Sorochinsky, A.E.; Fustero, S.; Soloshonok, V.A.; Liu, H. Fluorine in Pharmaceutical Industry: Fluorine-Containing Drugs Introduced to the Market in the Last Decade (2001–2011). Chem. Rev. 2014, 114, 2432–2506. [Google Scholar] [CrossRef]
  2. Inoue, M.; Sumii, Y.; Shibata, N. Contribution of Organofluorine Compounds to Pharmaceuticals. ACS Omega 2020, 5, 10633–10640. [Google Scholar] [CrossRef]
  3. Purser, S.; Moore, P.R.; Swallow, S.; Gouverneur, V. Fluorine in medicinal chemistry. Chem. Soc. Rev. 2008, 37, 320–330. [Google Scholar] [CrossRef]
  4. Zhou, Y.; Wang, J.; Gu, Z.; Wang, S.; Zhu, W.; Aceña, J.L.; Soloshonok, V.A.; Izawa, K.; Liu, H. Next Generation of Fluorine-Containing Pharmaceuticals, Compounds Currently in Phase II–III Clinical Trials of Major Pharmaceutical Companies: New Structural Trends and Therapeutic Areas. Chem. Rev. 2016, 116, 422–518. [Google Scholar] [CrossRef] [PubMed]
  5. Sedgwick, D.M.; Román, R.; Barrio, P.; Trabanco, A.A.; Fustero, S. Biorelevant fluorine-containing N-heterocycles. In Fluorine in Life Sciences: Pharmaceuticals, Medicinal Diagnostics, and Agrochemicals; Haufe, G., Leroux, F.R., Eds.; Academic Press: Cambridge, MA, USA, 2019; pp. 575–606. [Google Scholar]
  6. Zhang, B.; Wang, J. Assembly of versatile fluorine-containing structures via N-heterocyclic carbene organocatalysis. Sci. China Chem. 2022, 65, 1691–1703. [Google Scholar] [CrossRef]
  7. Mykhailiuk, P.K. Fluorinated Pyrazoles: From Synthesis to Applications. Chem. Rev. 2021, 121, 1670–1715. [Google Scholar] [CrossRef]
  8. Barata-Vallejo, S.; Cooke, M.V.; Postigo, A. Radical Fluoroalkylation Reactions. ACS Catal. 2018, 8, 7287–7307. [Google Scholar] [CrossRef]
  9. Baguia, H.; Evano, G. Direct Perfluoroalkylation of C−H Bonds in (Hetero) arenes. Chem.—Eur. J. 2022, 28, e202200975. [Google Scholar] [CrossRef] [PubMed]
  10. Hagmann, W.K. The Many Roles for Fluorine in Medicinal Chemistry. J. Med. Chem. 2008, 51, 4359–4369. [Google Scholar] [CrossRef]
  11. Barata-Vallejo, S.; Bonesi, S.M.; Postigo, A. Photocatalytic fluoroalkylation reactions of organic compounds. Org. Biomol. Chem. 2015, 13, 11153–11183. [Google Scholar] [CrossRef] [Green Version]
  12. Koike, T. Frontiers in Radical Fluoromethylation by Visible-Light Organic Photocatalysis. Asian J. Org. Chem. 2020, 9, 529–537. [Google Scholar] [CrossRef]
  13. Kliś, T. Visible-Light Photoredox Catalysis for the Synthesis of Fluorinated Aromatic Compounds. Catalysts 2023, 13, 94. [Google Scholar] [CrossRef]
  14. Gisbertz, S.; Pieber, B. Heterogeneous Photocatalysis in Organic Synthesis. ChemPhotoChem 2020, 4, 456–475. [Google Scholar] [CrossRef] [Green Version]
  15. Ghosh, I.; Khamrai, J.; Savateev, A.; Shlapakov, N.; Antonietti, M.; König, B. Organic semiconductor photocatalyst can bifunctionalize arenes and heteroarenes. Science 2019, 365, 360–366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Wang, L.; Liu, M.; Zha, W.; Wei, Y.; Ma, X.; Xu, C.; Lu, C.; Qin, N.; Gao, L.; Qiu, W.; et al. Mechanistic study of visible light-driven CdS or g-C3N4-catalyzed CH direct trifluoromethylation of (hetero) arenes using CF3SO2Na as the trifluoromethyl source. J. Catal. 2020, 389, 533–543. [Google Scholar] [CrossRef]
  17. Wen, J.; Xie, J.; Chen, X.; Li, X. A review on g-C3N4-based photocatalysts. Appl. Surf. Sci. 2017, 391, 72–123. [Google Scholar] [CrossRef]
  18. Singh, P.P.; Srivastava, V. Recent advances in visible-light graphitic carbon nitride (g-C3N4) photocatalysts for chemical transformations. RSC Adv. 2022, 12, 18245–18265. [Google Scholar] [CrossRef]
  19. Qin, H.-T.; Wu, S.-W.; Liu, J.-L.; Liu, F. Photoredox-catalysed redox-neutral trifluoromethylation of vinyl azides for the synthesis of α-trifluoromethylated ketones. Chem. Commun. 2017, 53, 1696–1699. [Google Scholar] [CrossRef]
  20. Wang, Y.-F.; Lonca, G.H.; Chiba, S. PhI(OAc)2-Mediated Radical Trifluoromethylation of Vinyl Azides with Me3SiCF3. Angew. Chem. Int. Ed. 2014, 53, 1067–1071. [Google Scholar] [CrossRef]
  21. Ikeda, I.; Kogame, Y.; Okahara, M. Synthesis of novel pyrazoles containing perfluoroalkyl groups by reactions of perfluoro-2-methylpent-2-ene and hydrazones. J. Org. Chem. 1985, 50, 3640–3642. [Google Scholar] [CrossRef]
  22. Paveliev, S.A.; Churakov, A.I.; Alimkhanova, L.S.; Segida, O.O.; Nikishin, G.I.; Terent’ev, A.O. Electrochemical Synthesis of O-Phthalimide Oximes from α-Azido Styrenes via Radical Sequence: Generation, Addition and Recombination of Imide-N-Oxyl and Iminyl Radicals with C−O/N−O Bonds Formation. Adv. Synth. Catal. 2020, 362, 3864–3871. [Google Scholar] [CrossRef]
  23. Sharma, P.; Sarngan, P.P.; Lakshmanan, A.; Sarkar, D. One-step synthesis of highly reactive g-C3N4. J. Mater. Sci. Mater. Electron. 2022, 33, 9116–9125. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of 3-(4-fluorophenyl)-1-(1-(4-fluorophenyl)-3,3,3-trifluoroprop-1-en-1-yl)-5-fluoro-1H-pyrazole 2.
Scheme 1. Synthesis of 3-(4-fluorophenyl)-1-(1-(4-fluorophenyl)-3,3,3-trifluoroprop-1-en-1-yl)-5-fluoro-1H-pyrazole 2.
Molbank 2023 m1620 sch001
Scheme 2. Synthetic transformation of vinyl azide 1 for preparation of fluorinated pyrazole 2.
Scheme 2. Synthetic transformation of vinyl azide 1 for preparation of fluorinated pyrazole 2.
Molbank 2023 m1620 sch002
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

Paveliev, S.A.; Ustyuzhanin, A.O.; Krylov, I.B.; Terent’ev, A.O. 3-(4-Fluorophenyl)-1-(1-(4-fluorophenyl)-3,3,3-trifluoroprop-1-en-1-yl)-5-fluoro-1H-pyrazole. Molbank 2023, 2023, M1620. https://doi.org/10.3390/M1620

AMA Style

Paveliev SA, Ustyuzhanin AO, Krylov IB, Terent’ev AO. 3-(4-Fluorophenyl)-1-(1-(4-fluorophenyl)-3,3,3-trifluoroprop-1-en-1-yl)-5-fluoro-1H-pyrazole. Molbank. 2023; 2023(2):M1620. https://doi.org/10.3390/M1620

Chicago/Turabian Style

Paveliev, Stanislav A., Alexander O. Ustyuzhanin, Igor B. Krylov, and Alexander O. Terent’ev. 2023. "3-(4-Fluorophenyl)-1-(1-(4-fluorophenyl)-3,3,3-trifluoroprop-1-en-1-yl)-5-fluoro-1H-pyrazole" Molbank 2023, no. 2: M1620. https://doi.org/10.3390/M1620

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