Next Article in Journal
Production of Prenylated Stilbenoids in Hairy Root Cultures of Peanut (Arachis hypogaea) and Its Wild Relatives A. ipaensis and A. duranensis via an Optimized Elicitation Procedure
Previous Article in Journal
From Dermal Patch to Implants—Applications of Biocomposites in Living Tissues
Previous Article in Special Issue
Precursor-Dependent Photocatalytic Activity of Carbon Dots
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Visible-light Promoted Atom Transfer Radical Addition−Elimination (ATRE) Reaction for the Synthesis of Fluoroalkylated Alkenes Using DMA as Electron-donor

1
Key Laboratory of Biocatalysis & Chiral Drug Synthesis of Guizhou Province, Generic Drug Research Center of Guizhou Province. School of Pharmacy, Zunyi Medical University, Zunyi 563003, China
2
Basic Medical School, Zunyi Medical University, Zunyi 563003, China
3
School of Medicine, Washington University in St. Louis, St. Louis, MO 63110, United States
4
Key Laboratory of Basic Pharmacology of Ministry of Education and Joint International Research Laboratory of Ethnomedicine of Ministry of Education, School of Pharmacy, Zunyi Medical University, Zunyi 563003, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2020, 25(3), 508; https://doi.org/10.3390/molecules25030508
Submission received: 12 December 2019 / Revised: 4 January 2020 / Accepted: 21 January 2020 / Published: 24 January 2020
(This article belongs to the Special Issue Photocatalytic Strategies in Organic Synthesis)

Abstract

:
Here, we describe a mild, catalyst-free and operationally-simple strategy for the direct fluoroalkylation of olefins driven by the photochemical activity of an electron donor−acceptor (EDA) complex between DMA and fluoroalkyl iodides. The significant advantages of this photochemical transformation are high efficiency, excellent functional group tolerance, and synthetic simplicity, thus providing a facile route for further application in pharmaceuticals and life sciences.

1. Introduction

Owing to their tendency to alter the lipophilicity, metabolic stability, and electronic properties of organic molecules, fluorinated organic compounds have been widely used in medicinal chemistry, and material sciences [1,2,3,4,5,6,7,8,9]. Therefore, the development of safer, less toxic and more selective methods to introduce fluorinated functional groups into organic molecules has become an intensive topic of synthetic organic chemistry [10,11,12,13,14,15,16,17,18,19].
Alkenes play a ubiquitous role in the realm of chemical synthesis due to their enriched reactivity and abundance, fluoroalkylation of carbon-carbon double bonds is an attractive method for accessing fluorine containing compounds. Since 1945, atom transfer radical addition (ATRA) has extensively been utilized for fluoroalkylation of alkenes [20,21,22,23,24,25,26,27,28,29]. However, the preparation of alkenes containing fluorinated functional groups via Heck-type reaction were less studied owning to the lack of efficient and general strategies [30,31,32]. In the past 10 years, some powerful strategies have been developed to realize such transformation [20,21,33,34,35,36]. One of the major improvements has been made via photo-excited catalyst such as Ru/Ir complexes and organic dyes [21,33]. Very recently, non-covalent interaction initiated fluoroalkylation reaction has emerged as an attractive strategy [37,38,39,40,41,42,43,44,45,46,47]. Inspired by our previous studies in this field [48,49,50,51], we envision that if the solvent can serve as an electron donor compound, the reactions would be simpler. Here, we demonstrate a mild, catalyst-free and operationally-simple strategy for the direct fluoroalkylation of olefins driven by the photochemical activity of electron donor−acceptor (EDA) complex between DMA and fluoroalkyl iodides. The significant advantages of this photochemical transformation are high efficiency, excellent functional group tolerance, and synthetic simplicity [52].

2. Results

We initially probed this catalyst-free fluoroalkylation reaction by using readily available tert-butyl allylcarbamate 1a and ethyl iododifluoroacetate 2a (1.5 equiv) as model substrates. 29% yield of atom transfer radical addition (ATRA) product 4a was obtained when the reaction was performed with K3PO4 (2.0 equiv) in MeCN and irradiated by blue LEDs for 16 h. After a series of reaction media were screened (Table 1, entries 2–8), THF, Toluene and dioxane were not suitable for this transformation, and 10% yield of desired product was obtained when DMSO was used as solvent. Both 3a and 4a were obtained when DMA and DMF were used. To improve the selectivity of this transformation, a variety of different bases were examined utilizing DMA as solvent for this reaction (Table 1, entries 8–11). Among them, KOAc was the best choice and afforded 3a in 64% yield (Table 1, entry 11). Finally, different light sources were tested, the yield increased to 95% when the reaction mixtures were irradiated under purple LEDs (Table 1, entry 13), no desired product was observed when control experiments were carried out in the absence of light or base, demonstrating the photochemical nature of this transformation (Table 1, entries 14, 15).
With the optimum reaction conditions established, various alkenes were explored. As shown in Scheme 1, The reaction exhibited good functional group tolerance. A range of functional groups, such as esters (3c), methoxyl (3d), and even unprotected hydroxyl group (3e) generally were compatible with the reaction and moderate to good yields were obtained. Then, styrenes were also evaluated, and we were pleased to find that styrenes all performed well under the optimized reaction conditions, providing the corresponding products in good to excellent yields (3fn). Importantly, substrate with steric hindrance could also undergo this transformation smoothly (3o). This reaction system is also amenable to the use of other commercial perfluoroalkyl iodides, such as C4F9I, C6F13I and C8F17I (Scheme 1, 3ps).
To gain insight into the mechanism of this reaction, several experiments were performed. The reaction was completely suppressed when a radical scavenger TEMPO (1.0 equiv) was added as an additive under the standard reaction conditions (Scheme 2a). Furthermore, a radical clock experiment was conducted. The ring-expanded product 7 was formed when 2a was treated with α-cyclopropylstyrene (6) in the absence of 1a (Scheme 2b). Optical absorption spectra of the reactants found that the absorption was obvious strengthened when DMA and RFI were mixed (Scheme 2c, for details, see Supplementary Materials), which indicated a non-covalent interaction occurred between them. Moreover, this conclusion was further confirmed by a Job’s plot (Scheme 2d, for details, see Supplementary Materials). Finally, compound 4a was totally converted to 3a under optimized condition in the dark (Scheme 2e, for details, see Supplementary Materials). This result indicated that the reaction was performed via atom transfer radical addition elimination pathway.
Based on these preliminary results and previous reports [53,54], a plausible mechanism was depicted in Scheme 3. Initially, non-covalent interactions occurred between DMA and C–I bond. Then fluoroalkyl radical was generated under the irradition of purple LEDs. Subsequently, fluoroalkyl radical reacted with alkenes (1) and generated a carbon radical A, which abstracted an iodine atom from RFI to afford ATRA product 4 along with RF radical that sustains the chain. Finally, elimination of 4 with base could afford the corresponding fluoroalkylated-ATRE products (3).

3. Discussion

In the past 10 years, fluoroalkylation involving visible light promoted reactions has emerged as an attractive and useful strategy to directly introduce fluorinated groups into organic molecules Among those methods, fluoroalkyl iodides were commonly used substrates. Based on the investigations in our group, we believe the noncovalent interaction between solvents (DMF, DMA, MeCN, acetone ect) and fluoroalkyl iodides might play a crucial role in this kind of transformation.

4. Materials and Methods

1H NMR spectra and 13C NMR spectra were recorded on an Agilent AM400 spectrometer. 19F NMR was recorded on an Agilent 400 MHz NMR spectrometer (CFCl3 as outside standard and low field is positive). Chemical shifts (δ) are reported in ppm, and coupling constants (J) are in Hertz (Hz). The following abbreviations were used to explain the multiplicities: s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet, br = broad. NMR yield was determined by 19F NMR using fluorobenzene as an internal standard before working up the reaction.
All reagents were used as received from commercial sources, unless specified otherwise, or prepared as described in the literature. All reagents were weighed and handled in air at room temperature. Blue LEDs (430–490 nm, peak wavelength: 455.0 nm) and purple LEDs (380–425 nm, peak wavelength: 395.0 nm) were bought online.

Supplementary Materials

The supplementary materials are available online.

Author Contributions

C.-Y.H. and X.-F.L. conceived and designed the experiments. W.-W.X., L.W. and T.M. performed the experiments and mechanism studies. C.-Y.H. and X.-F.L. co-wrote the manuscript. J.G. contributed for the edition of manuscript at revision stage. All authors discussed the results and commented on the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

We are grateful for the financial support from the National Natural Science Foundation of China (No. 21702241, 81760624, 21762053), Programs of Guizhou Province (No. 2017-1225, 2018-1427), Guizhou Provincial Science & Technology “125-Plan” Major Project (No. 2015-039).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Purser, S.; Moore, P.R.; Swallow, S.; Gouverneur, V. Fluorine in medicinal chemistry. Chem. Soc. Rev. 2008, 37, 320–330. [Google Scholar] [CrossRef]
  2. Müller, K.; Faeh, C.; Diederich, F. Fluorine in Pharmaceuticals: Looking Beyond Intuition. Science 2007, 317, 1881–1886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. 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] [PubMed]
  4. Preshlock, S.; Tredwell, M.; Gouverneur, V. 18F-Labeling of Arenes and Heteroarenes for Applications in Positron Emission Tomography. Chem. Rev. 2016, 116, 719–766. [Google Scholar] [CrossRef] [PubMed]
  5. O’Hagan, D. Understanding organofluorine chemistry. An Introduction to the C–F bond. Chem. Soc. Rev. 2008, 37, 308–319. [Google Scholar] [CrossRef] [PubMed]
  6. Burgey, C.S.; Robinson, K.A.; Lyle, T.A.; Sanderson, P.E.J.; Dale Lewis, S.; Lucas, B.J.; Krueger, J.A.; Singh, R.; Miller-Stein, C.; White, R.B.; et al. Metabolism-Directed Optimization of 3-Aminopyrazinone Acetamide Thrombin Inhibitors. Development of an Orally Bioavailable Series Containing P1 and P3 Pyridines. J. Med. Chem. 2003, 46, 461–473. [Google Scholar] [CrossRef] [PubMed]
  7. Grunewald, G.L.; Seim, M.R.; Lu, J.; Makboul, M.; Criscione, K.R. Application of the Goldilocks Effect to the Design of Potent and Selective Inhibitors of Phenylethanolamine N-Methyltransferase:  Balancing pKa and Steric Effects in the Optimization of 3-Methyl-1,2,3,4-tetrahydroisoquinoline Inhibitors by β-Fluorination. J. Med. Chem. 2006, 49, 2939–2952. [Google Scholar] [CrossRef] [Green Version]
  8. Meanwell, N.A. Synopsis of Some Recent Tactical Application of Bioisosteres in Drug Design. J. Med. Chem. 2011, 54, 2529–2591. [Google Scholar] [CrossRef]
  9. Charpentier, J.; Früh, N.; Togni, A. Electrophilic Trifluoromethylation by Use of Hypervalent Iodine Reagents. Chem. Rev. 2015, 115, 650–682. [Google Scholar] [CrossRef]
  10. Alonso, C.; Marigorta, E.M.; Rubiales, G.; Palacios, F. Carbon Trifluoromethylation Reactions of Hydrocarbon Derivatives and Heteroarenes. Chem. Rev. 2015, 115, 1847–1935. [Google Scholar] [CrossRef]
  11. Tomashenko, O.A.; Grushin, V.V. Aromatic Trifluoromethylation with Metal Complexes. Chem. Rev. 2011, 111, 4475–4521. [Google Scholar] [CrossRef] [PubMed]
  12. Chu, L.; Qing, F.-L. Oxidative Trifluoromethylation and Trifluoromethylthiolation Reactions Using (Trifluoromethyl)trimethylsilane as a Nucleophilic CF3 Source. Acc. Chem. Res. 2014, 47, 1513–1522. [Google Scholar] [CrossRef]
  13. Xu, X.-H.; Matsuzaki, K.; Shibata, N. Synthetic Methods for Compounds Having CF3–S Units on Carbon by Trifluoromethylation, Trifluoromethylthiolation, Triflylation, and Related Reactions. Chem. Rev. 2015, 115, 731–764. [Google Scholar] [CrossRef]
  14. Feng, Z.; Xiao, Y.-L.; Zhang, X. Transition-Metal (Cu, Pd, Ni)-Catalyzed Difluoroalkylation via Cross-Coupling with Difluoroalkyl Halides. Acc. Chem. Res. 2018, 51, 2264–2278. [Google Scholar] [CrossRef] [PubMed]
  15. Chatterjee, T.; Iqbal, N.; You, Y.; Cho, E.J. Controlled Fluoroalkylation Reactions by Visible-Light Photoredox Catalysis. Acc. Chem. Res. 2016, 49, 2284–2294. [Google Scholar] [CrossRef] [PubMed]
  16. Yerien, D.E.; Barata-Vallejo, S.; Postigo, A. Difluoromethylation Reactions of Organic Compounds. Chem. Eur. J. 2017, 23, 14676–14701. [Google Scholar] [CrossRef] [PubMed]
  17. Moschner, J.; Stulberg, V.; Fernandes, R.; Huhmann, S.; Leppkes, J.; Koksch, B. Approaches to Obtaining Fluorinated α-Amino Acids. Chem. Rev. 2019, 119, 10718–10801. [Google Scholar] [CrossRef] [PubMed]
  18. Shao, X.; Xu, C.; Lu, L.; Shen, Q. Shelf-Stable Electrophilic Reagents for Trifluoromethylthiolation. Acc. Chem. Res. 2015, 48, 1227–1236. [Google Scholar] [CrossRef]
  19. Furuya, T.; Kamlet, A.S.; Ritter, T. Catalysis for Fluorination and Trifluoromethylation. Nature 2011, 473, 470–477. [Google Scholar] [CrossRef] [Green Version]
  20. Yajima, T.; Ikegami, M. Metal-Free Visible-Light Radical Iodoperfluoroalkylation of Terminal Alkenes and Alkynes. Eur. J. Org. Chem. 2017, 2017, 2126–2129. [Google Scholar] [CrossRef] [Green Version]
  21. Yu, C.; Iqbal, N.; Park, S.; Cho, E.J. Selective Difluoroalkylation of Alkenes by using Visible Light Photoredox Catalysis. Chem. Commun. 2014, 50, 12884–12887. [Google Scholar] [CrossRef]
  22. Tang, X.-J.; Dolbier, W.R. Efficient Cu-catalyzed Atom Transfer Radical Addition Reactions of Fluoroalkylsulfonyl Chlorides with Electron-deficient Alkenes Induced by Visible Light. Angew. Chem. Int. Ed. 2015, 54, 4246–4249. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, T.; Cheung, C.W.; Hu, X. Iron-Catalyzed 1,2-Addition of Perfluoroalkyl Iodides to Alkynes and Alkene. Angew. Chem. Int. Ed. 2014, 53, 4910–4914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Behrends, I.; Baehr, S.; Czekelius, C. Perfluoroalkylation of Alkenes by Frustrated Lewis Pairs. Chem. Eur. J. 2016, 22, 17177–17181. [Google Scholar] [CrossRef] [PubMed]
  25. Nguyen, J.D.; Tucker, J.W.; Konieczynska, M.D.; Stephenson, C.R.J. Intermolecular Atom Transfer Radical Addition to Olefins Mediated by Oxidative Quenching of Photoredox Catalysts. J. Am. Chem. Soc. 2011, 133, 4160–4163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Wu, G.; von Wangelin, J.A. Stereoselective Cobalt-catalyzed Halofluoroalkylation of Alkynes. Chem. Sci. 2018, 9, 1795–1802. [Google Scholar] [CrossRef] [Green Version]
  27. Beniazza, R.; Remisse, L.; Jardel, D.; Lastecoueres, D.; Vincent, J.-M. Light-mediated Iodoperfluoroalkylation of Alkenes/Alkynes Catalyzed by Chloride Ions: Role of Halogen Bonding. Chem. Commun. 2018, 54, 7451–7454. [Google Scholar] [CrossRef]
  28. Rawner, T.; Lutsker, E.; Kaiser, C.A.; Reiser, O. The Different Faces of Photoredox Catalysts: Visible-Light-Mediated Atom Transfer Radical Addition (ATRA) Reactions of Perfluoroalkyl Iodides with Styrenes and Phenylacetylenes. ACS Catal. 2018, 8, 3950–3956. [Google Scholar] [CrossRef]
  29. Magagnano, G.; Gualandi, A.; Marchini, M.; Mengozzi, L.; Ceroni, P.; Cozzi, P.G. Photocatalytic ATRA Reaction Promoted by Iodo-Bodipy and Sodium Ascorbate. Chem. Commun. 2017, 53, 1591–1594. [Google Scholar] [CrossRef]
  30. Long, Z.-Y.; Chen, Q.-Y. The Activation of Carbon−Chlorine Bonds in Per- and Polyfluoroalkyl Chlorides:  DMSO-Induced Hydroperfluoroalkylation of Alkenes and Alkynes with Sodium Dithionite. J. Org. Chem. 1999, 64, 4775–4782. [Google Scholar] [CrossRef]
  31. Murakami, S.; Ishii, H.; Fuchigami, T. Electrosynthesis of Ethyl a,a-difluoro-a-(phenylseleno)acetate and its Photochemical Synthetic Application. J. Fluorine. Chem. 2004, 125, 609–614. [Google Scholar] [CrossRef]
  32. Ghattas, W.; Hess, C.R.; Iacazio, G.; Hardre, R.; Klinman, J.P.; Reglier, M. Pathway for the Stereocontrolled Z and E Production of α,α-Difluorine-Substituted Phenyl Butenoates. J. Org. Chem. 2006, 71, 8618–8621. [Google Scholar] [CrossRef] [PubMed]
  33. Tiwari, D.P.; Dabral, S.; Wen, J.; Wiesenthal, J.; Terhorst, S.; Bolm, C. Organic Dye-Catalyzed Atom Transfer Radical Addition–Elimination (ATRE) Reaction for the Synthesis of Perfluoroalkylated Alkenes. Org. Lett. 2017, 19, 4295–4298. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, X.; Zhao, S.; Liu, J.; Zhu, D.; Guo, M.; Tang, X.; Wang, G. Copper-Catalyzed C–H Difluoroalkylations and Perfluoroalkylations of Alkenes and (Hetero)arenes. Org. Lett. 2017, 19, 4187–4190. [Google Scholar] [CrossRef] [PubMed]
  35. Feng, Z.; Min, Q.-Q.; Zhao, H.-Y.; Gu, J.-W.; Zhang, X. A General Synthesis of Fluoroalkylated Alkenes by Palladium-Catalyzed Heck-Type Reaction of Fluoroalkyl Bromides. Angew. Chem. Int. Ed. 2015, 54, 1270–1274. [Google Scholar] [CrossRef]
  36. Xie, J.; Li, J.; Weingand, V.; Rudolph, M.; Hashmi, A.S.K. Intermolecular Photocatalyzed Heck-like Coupling of Unactivated Alkyl Bromides by a Dinuclear Gold Complex. Chem. Eur. J. 2016, 22, 12646–12650. [Google Scholar] [CrossRef]
  37. Nappi, M.; Bergonzini, G.; Melchiorre, P. Metal-Free Photochemical Aromatic Perfluoroalkylation of α-Cyano Arylacetates. Angew. Chem. Int. Ed. 2014, 53, 4921–4925. [Google Scholar] [CrossRef]
  38. Fernández-Alvarez, V.M.; Nappi, M.; Melchiorre, P.; Maseras, F. Computational Study with DFT and Kinetic Models on the Mechanism of Photoinitiated Aromatic Perfluoroalkylations. Org. Lett. 2015, 17, 2676–2679. [Google Scholar] [CrossRef]
  39. Kandukuri, S.R.; Bahamonde, A.; Chatterjee, I.; Jurberg, I.D.; Escudero-Adán, E.C.; Melchiorre, P. X-Ray Characterization of an Electron Donor–Acceptor Complex that Drives the Photochemical Alkylation of Indoles. Angew. Chem. Int. Ed. 2015, 54, 1485–1489. [Google Scholar] [CrossRef]
  40. Guo, Q.; Wang, M.; Liu, H.; Wang, R.; Xu, Z. Visible-Light-Promoted Dearomative Fluoroalkylation of β-Naphthols through Intermolecular Charge Transfer. Angew. Chem. Int. Ed. 2018, 57, 4747–4751. [Google Scholar] [CrossRef]
  41. Wang, Y.; Wang, J.; Li, G.-X.; He, G.; Chen, G. Halogen-Bond-Promoted Photoactivation of Perfluoroalkyl Iodides: A Photochemical Protocol for Perfluoroalkylation Reactions. Org. Lett. 2017, 19, 1442–1445. [Google Scholar] [CrossRef] [PubMed]
  42. Helmecke, L.; Spittler, L.; Baumgarten, K.; Czekelius, C. Metal-Free Activation of C–I Bonds and Perfluoroalkylation of Alkenes with Visible Light Using Phosphine Catalysts. Org. Lett. 2019, 21, 7823–7827. [Google Scholar] [CrossRef] [PubMed]
  43. Liu, Y.; Chen, X.-L.; Sun, K.; Li, X.-Y.; Zeng, F.-L.; Liu, X.-C.; Qu, L.-B.; Zhao, Y.-F.; Yu, B. Visible-Light Induced Radical Perfluoroalkylation/Cyclization Strategy To Access 2-Perfluoroalkylbenzothiazoles/Benzoselenazoles by EDA Complex. Org. Lett. 2019, 21, 4019–4024. [Google Scholar] [CrossRef] [PubMed]
  44. Tu, H.-Y.; Zhu, S.; Qing, F.-L.; Chu, L. A Four-component Radical Cascade Trifluoromethylation Reaction of Alkenes Enabled by an Electron-donor–acceptor Complex. Chem. Commun. 2018, 54, 12710–12713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Park, G.-R.; Choi, Y.; Choi, M.G.; Chang, S.-K.; Cho, E.J. Metal-Free Visible-Light-Induced Trifluoromethylation Reactions. Asian J. Org. Chem. 2017, 6, 436–440. [Google Scholar] [CrossRef]
  46. Cheng, Y.; Yuan, X.; Ma, J.; Yu, S. Direct Aromatic C-H Trifluoromethylation via an Electron-Donor–Acceptor Complex. Chem. Eur. J. 2015, 21, 8355–8359. [Google Scholar] [CrossRef]
  47. Postigo, A. Electron Donor-Acceptor Complexes in Perfluoroalkylation Reactions. Eur. J. Org. Chem. 2018, 46, 6391–6404. [Google Scholar] [CrossRef]
  48. Zhao, L.; Huang, Y.; Wang, Z.; Zhu, E.; Mao, T.; Jia, J.; Gu, J.; Li, X.-F.; He, C.-Y. Organophosphine-Catalyzed Difluoroalkylation of Alkenes. Org. Lett. 2019, 21, 6705–6709. [Google Scholar] [CrossRef]
  49. Zhu, E.; Liu, X.-X.; Wang, A.-J.; Mao, T.; Zhao, L.; Zhang, X.; He, C.-Y. Visible Light Promoted Fluoroalkylation of Alkenes and Alkynes using 2-Bromophenol as a Catalyst. Chem. Commun. 2019, 55, 12259–12262. [Google Scholar] [CrossRef]
  50. Huang, Y.; Lei, Y.-Y.; Zhao, L.; Gu, J.; Yao, Q.; Wang, Z.; Li, X.-F.; Zhang, X.; He, C.-Y. Catalyst-free and Visible Light Promoted Trifluoromethylation and Perfluoroalkylation of Uracils and Cytosines. Chem. Commun. 2018, 54, 13662–13665. [Google Scholar] [CrossRef]
  51. Mao, T.; Ma, M.-J.; Zhao, L.; Xue, D.-P.; Yu, Y.; Gu, J.; He, C.-Y. A General and Green Fluoroalkylation Reaction Promoted via Noncovalent Interactions between Acetone and Fluoroalkyl Iodides. Chem. Commun. 2020, 56. [Google Scholar] [CrossRef] [PubMed]
  52. Li, K.; Zhang, X.; Chen, J.; Gao, Y.; Yang, C.; Zhang, K.; Zhou, Y.; Fan, B. Blue Light Induced Difluoroalkylation of Alkynes and Alkenes. Org. Lett. 2019, 21, 9914–9918. [Google Scholar] [CrossRef] [PubMed]
  53. Cismesia, M.; Yoon, T. Characterizing Chain Processes in Visible Light Photoredox Catalysis. Chem. Sci. 2015, 6, 5426–5434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Buzzetti, L.; Crisenza, G.E.; Melchiorre, P. Mechanistic Studies in Photocatalysis. Angew. Chem. Int. Ed. 2019, 58, 3730–3747. [Google Scholar] [CrossRef] [Green Version]
Sample Availability: Not available.
Scheme 1. Atom Transfer Radical Addition−Elimination (ATRE) reaction for the synthesis of fluoroalkylated alkenes a,b. a Reaction conditions (unless otherwise specified): 1 (0.3 mmol, 1.0 equiv), 2 (0.45 mmol, 1.5 equiv), KOAc (0.6 mmol, 2.0 equiv) in anhydrous DMA (2.0 mL), r.t. under Ar, purple LEDs, for 16 h. b Yield of isolated product.
Scheme 1. Atom Transfer Radical Addition−Elimination (ATRE) reaction for the synthesis of fluoroalkylated alkenes a,b. a Reaction conditions (unless otherwise specified): 1 (0.3 mmol, 1.0 equiv), 2 (0.45 mmol, 1.5 equiv), KOAc (0.6 mmol, 2.0 equiv) in anhydrous DMA (2.0 mL), r.t. under Ar, purple LEDs, for 16 h. b Yield of isolated product.
Molecules 25 00508 sch001
Scheme 2. Mechanistic investigation. (a) Addition of radical and SET inhibitors. (b) Trapping of intermediates. (c) Optical absorption spectra study. (d) Job’s plot. (e) Control experiment.
Scheme 2. Mechanistic investigation. (a) Addition of radical and SET inhibitors. (b) Trapping of intermediates. (c) Optical absorption spectra study. (d) Job’s plot. (e) Control experiment.
Molecules 25 00508 sch002
Scheme 3. Proposed Reaction Mechanism.
Scheme 3. Proposed Reaction Mechanism.
Molecules 25 00508 sch003
Table 1. Representative results for optimization of visible light-mediated reaction of 1a and 2a. a,bMolecules 25 00508 i001
Table 1. Representative results for optimization of visible light-mediated reaction of 1a and 2a. a,bMolecules 25 00508 i001
EntryLight SourceBase (equiv)Sovent3a, yield (%)4a, yield (%)
1Blue LEDsK3PO4 (2)MeCN----29
2Blue LEDsK3PO4 (2)DCE----55
3Blue LEDsK3PO4 (2)THF--------
4Blue LEDsK3PO4 (2)Toluene----7
5Blue LEDsK3PO4 (2)Dioxane----trace
6Blue LEDsK3PO4 (2)DMSO10----
7Blue LEDsK3PO4 (2)DMF2151
8Blue LEDsK3PO4 (2)DMA4536
9Blue LEDsK2CO3 (2)DMA4326
10Blue LEDsCs2CO3 (2)DMA595
11Blue LEDsKOAc (2)DMA64----
12Green LEDsKOAc (2)DMA50----
13Purple LEDsKOAc (2)DMA95(84)----
14Purple LEDsNoneDMA--------
15 cnoneKOAc (2)DMA--------
a Reaction conditions: (unless otherwise specified): 1a (0.3 mmol, 1.0 equiv), 2a (0.45 mmol, 1.5 equiv), anhydrous solvent (2 mL), r.t. under Ar, irritated under visible light for 16 h. b Determined by 19F NMR spectroscopy using fluorobenzene as an internal standard, and the number within parentheses represents the yield of the isolated product. c The reaction was performed without light.

Share and Cite

MDPI and ACS Style

Xu, W.-W.; Wang, L.; Mao, T.; Gu, J.; Li, X.-F.; He, C.-Y. Visible-light Promoted Atom Transfer Radical Addition−Elimination (ATRE) Reaction for the Synthesis of Fluoroalkylated Alkenes Using DMA as Electron-donor. Molecules 2020, 25, 508. https://doi.org/10.3390/molecules25030508

AMA Style

Xu W-W, Wang L, Mao T, Gu J, Li X-F, He C-Y. Visible-light Promoted Atom Transfer Radical Addition−Elimination (ATRE) Reaction for the Synthesis of Fluoroalkylated Alkenes Using DMA as Electron-donor. Molecules. 2020; 25(3):508. https://doi.org/10.3390/molecules25030508

Chicago/Turabian Style

Xu, Wen-Wen, Le Wang, Ting Mao, Jiwei Gu, Xiao-Fei Li, and Chun-Yang He. 2020. "Visible-light Promoted Atom Transfer Radical Addition−Elimination (ATRE) Reaction for the Synthesis of Fluoroalkylated Alkenes Using DMA as Electron-donor" Molecules 25, no. 3: 508. https://doi.org/10.3390/molecules25030508

Article Metrics

Back to TopTop