Abstract
Direct C–H fluorination is an efficient strategy to construct aromatic C–F bonds, but the cleavage of specific C–H bonds in the presence of other functional groups and the high barrier of C–F bond formation make the transformation challenging. Progress for the electrophilic fluorination of arenes has been reported, but a similar transformation for electron-deficient azaarenes has remained elusive due to the high energy of the corresponding Wheland intermediates. Nucleophilic fluorination of electron-deficient azaarenes is difficult owing to the identity of the Meisenheimer intermediate after fluoride attack, from which fluoride elimination to regenerate the substrate is favored over hydride elimination to form the product. Herein, we report a new concept for C–H nucleophilic fluorination without the formation of azaarene Meisenheimer intermediates through a chain process with an asynchronous concerted F–-e–-H+ transfer. The concerted nucleophilic aromatic substitution strategy allows for the first successful nucleophilic oxidative fluorination of quinolines.
Single-step transformations from aromatic C–H bonds to C–F bonds are attractive methodologies because no prefunctionalization is required to obtain valuable fluorinated products, which often exhibit desirable properties and are widely applied in medicinal and materials chemistry.1−3 However, selective aromatic C–H bond fluorination of functionalized molecules is far from being established, in part due to a high reaction barrier associated with C–F bond formation that results from the high electronegativity of fluorine and the small ionic radius of fluoride (1.33 Å),4 both of which make reductive elimination from metal fluorides difficult. Therefore, aromatic C–H fluorination is a comparatively underdeveloped area with just a few successful approaches so far.5−11 Fluorination of six-membered azaarenes is even more challenging, and there is currently no method for a C4-selective fluorination at all.12,13 The sp2-hybridized nitrogen atom on electron-deficient azaarenes renders electrophilic aromatic substitution (SEAr) difficult due to a highly unstable potential Wheland intermediate (Figure 1a).6,13 Radical aromatic substitution of azaarenes such as Minisci-type reactions can afford C–C, C–B and C–Si bonds14−18 but not C–F bonds.19 Nucleophilic fluorination is challenging because, after fluoride addition to form a Meisenheimer intermediate, a hydride must be removed for rearomatization. Moreover, formation of the fluoride Meisenheimer intermediate is endergonic for azaarenes (Figure 1b), which would require a facile hydride elimination so that the overall barrier is not prohibitively large. Elimination of hydride from the fluoride Meisenheimer intermediate is challenging because the high-lying HOMO upon fluoride attack has overlap with the low-lying antibonding orbital of the C–F bond (σ*C–F), which results in the weakening of the C–F bond and facilitates heterolytic cleavage back to the starting material (Figure 1b, bottom). Modern nucleophilic aromatic substitution (SNAr) strategies to selectively attach a linchpin such as a phosphonium substituent at the C4 position of azaarenes for subsequent functionalization have found broad applications in C–H functionalization, but the strategy has not yet been extended to fluorination.20,21 So far, only a C2-selective fluorination was realized through a AgF2-mediated Chichibabin-type reaction (Figure 1c).10 While the C4 position is the most electrophilic site of six-membered azaarenes,12,13 C–H fluorination at this position has not been accomplished.22−26 Here we report a new approach to overcome the fundamental challenge for cleaving the strong C–H bond without the formation of high-energy Meisenheimer intermediates and successfully apply the concept to the C–H fluorination of quinolines (Figure 1d).
Figure 1.
(a) High-energy Wheland intermediate in the SEAr mechanism. (b) High-energy fluoride Meisenheimer intermediate in the SNAr mechanism. (c) C2-selective fluorination of azaarenes with AgF2. (d) This work: fluorination of azaarenes via concerted nucleophilic substitution. M, Meisenheimer complex; TS, transition state.
Meisenheimer complexes are commonly considered intermediates in SNAr reactions. However, in concerted nucleophilic aromatic substitutions (CSNAr),24,27,28 Meisenheimer complexes are transition states, in which case the formation of high-energy intermediates is avoided. Therefore, a CSNAr fluorination may provide the opportunity to avoid the high barriers that are associated with elimination of hydride from an already high-energy Meisenheimer complex to achieve azaarene C–H fluorination. We envisioned that when the fluoride attack on a protonated azaarene was coupled to an electron transfer, C–H bond heterolytic cleavage of a radical cation to release H+ would be more facile and allow a concerted process. Whether presumed F–, e–, and H+ transfer are concerted depends on the driving forces of both electron transfer (ET) and proton transfer (PT) as well as the distance between the fluoride Meisenheimer complexes and e– and H+ acceptors.29,30 For example, preassociation of reactants can play an important role in concerted proton-coupled electron transfer.29,30 Therefore, we attempted to generate an ion pair TEDA2+•F– (TEDA, N-(chloromethyl)triethylenediamine)31−35 in proximity to the protonated quinoline for potential concerted F–-e– transfer, simultaneous with deprotonation (Figure 2a, chain propagation). We have previously made use of the high electron affinity of the doubly cationic radical TEDA2+• (12.4 eV)31 for a charge-transfer-directed radical substitution. We therefore envisioned that TEDA2+•F– may support an F–-e–-H+ transfer (Figure 2a, A → D) and then e– transfer (D → A) chain for C–H fluorination upon further single-electron oxidation of the intermediate after F– attack and C–H cleavage. In this case, the formation of the protonated Meisenheimer intermediate could be avoided, and the challenging fluorination could become feasible. The overall two-electron reaction Ar-H + Selectfluor → Ar-F + TEDA-H2+ would thus be achieved via two single-electron redox processes, and ET to sustain the chain process.
Figure 2.
(a) Nucleophilic fluorination of azaarene via a radical chain mechanism; ΔGsolv298K in acetonitrile.42 (b) Representative fluorination reaction of quinoline. PS, photosensitizer; BET, back electron transfer; and PT, proton transfer.
Chain initiation from Selectfluor by single-electron reduction to yield the TEDA2+•F– ion pair for chain propagation requires an electron donor that fulfills the following criteria: the electron donor is strong enough to reduce Selectfluor under acidic conditions for a sufficiently high concentration of TEDA2+•; and the electron donor should not engage in other side reactions with Selectfluor. An appropriate species that meets both requirements could be the reduced state of protonated azaarene itself, the N-heterocyclic π-radical F(36,37) (Figure 2a, chain initiation). Our group has reported that the excitation of an ion pair acridine-H+Cl– followed by reductive quenching of its counteranion Cl– can produce a N-heterocyclic π-radical,38 and we intended to apply the Cl– quenching process to excited-state protonated azaarene E to generate a π-radical F. The electron donor F can thus donate an electron to Selectfluor to generate TEDA2+•F– for chain propagation. The excited-state protonated azaarene E could originate from an energy transfer (EnT)39 between the ground-state protonated azaarene and a triplet-state photosensitizer (PS). For EnT to occur effectively, we chose xanthone as a photosensitizer due to its high triplet-state energy of ET = 73.8 kcal/mol40 and a quinoline derivative as the substrate (ET = 57.7 kcal/mol41 for protonated quinoline) to evaluate the fluorination of 1. When HCl and Et3N·HCl were used as H+ and Cl– donors, respectively, C4-fluorinated quinoline derivative 2 was obtained in 56% isolated yield upon irradiation with a 365 nm LED (Figure 2b).
Concerted Mechanism
Upon initiation to generate TEDA2+•F–, fluoride-coupled electron transfer to protonated quinoline A, simultaneous with deprotonation, to generate D is associated with an activation energy of 18.3 kcal/mol (Figure 2a, chain propagation). The electron-rich π-radical D can be oxidized by Selectfluor to form the fluorinated product and regenerate TEDA2+•F– to complete the chain, which is exergonic by 46.2 kcal/mol.
Stepwise Mechanism 1
The fluoride of TEDA2+•F– could attack quinoline hydrochloride A to form Meisenheimer intermediate B and TEDA2+•, followed by ET-PT or HAT of B (A → B → D). Estimation of the ET activation energy for the reaction between B and TEDA2+• shows that there is no significant barrier (Figure 3a, ΔG⧧ < 0.1 kcal/mol), which indicates that it is unlikely for the Meisenheimer intermediate B and TEDA2+• to form a stable complex. An internal reaction coordinate (IRC) analysis revealed that the transition state of TEDA2+•F– attacking quinoline hydrochloride A directly leads to the product after ET, which excludes the existence of additional maxima along the reaction path (Figure 3b). Natural bond orbital analysis (NBO) of the transition state indicates a significant charge transfer process during the C–F bond formation, suggesting that an ET is coupled with the fluoride attack (Figure 3b). The calculated HOMO (Figure 3a, bottom) of intermediate B shows that the fluoride and enamine parts have the major contribution to the HOMO, which is remote from the proton donor. The lack of HOMO contribution from the C–H bond indicates that the C–H cleavage is difficult for a traditional HAT pathway,43,44 which would require the H+ and e– to originate from the same donor group.29,30,45
Figure 3.
Mechanistic investigation. (a) Estimation of the ΔG⧧SET. DFT-calculated geometry and HOMO of B; isosurface value = 0.06. (b) Reaction coordinate of the fluoride-coupled electron transfer. (c) KIE from an intramolecular competition reaction. (d) Reaction coordinate of the PT. (e) Relationship between the yield of fluorination and ET of photosensitizers. (f) Effect of chloride and a chlorine-radical-trapping reaction.
Stepwise Mechanism 2
A fluoride-coupled electron transfer mechanism to form a complex of dihydroquinoline radical cation46,47C and TEDA+, followed by a second step of deprotonation (A → C → D), could be feasible as well. Yet, we determined a kinetic isotope effect (KIE) of 1.7 from an intermolecular competition experiment with quinoline and quinoline-d7 (Figure 3c), which indicates a PT equilibrium before the product-formation ET (D → ArF), which is consistent with a concerted F–-e–-H+ mechanism (A → D) or a stepwise mechanism (A → C → D) with a reversible deprotonation. From the calculation, when Cl– or TEDA+ were used as proton acceptors, the deprotonation proceeds without a significant energy barrier (Figure 3d, AfterTS, ΔG⧧ < 0.1 kcal/mol), which is consistent with an asynchronous concerted F–-e–-H+ transfer mechanism48 from A to D.
Chain Initiation
The fluorination reaction efficiency correlates with the ET of the photosensitizers but not with their reduction potentials, which is in agreement with an EnT process to form the triplet state of protonated quinoline (Figure 3e).39 Control experiments show the important role of the Cl– counteranion and the Et3N·HCl additive to improve the reaction yield (Figure 3f), which could indicate a further reductive quenching process of the triplet state of protonated quinoline by Cl–. The observation of 1-chlorododecane in an intermolecular radical-trapping experiment with 1-dodecene is consistent with the formation of Cl·.38,49 According to an electrochemistry study, the additive Et3N·HCl (Ep1/2 = +0.94 V vs SCE) cannot be oxidized by ground-state protonated quinoline (Ep1/2 = −1.03 V vs SCE). However, the reduction of excited-state protonated quinoline (E0–0 = 2.65 V,50Ep1/2* = +1.62 V vs SCE) by Et3N·HCl is feasible. The quantum yield for fluorination of 3.5% is consistent with a back ET (BET) between N-heterocyclic π-radical F and Cl·.
The fluorination reaction can be applied to small molecules with quinoline scaffolds, as shown in Figure 4. The conditions enable C–H fluorination in the presence of a range of functional groups, including esters (7, 8, 13), halogens (9, 12, 17), ketone (4), cyano (16), phosphoryl (18), alkyls (5, 7), fluoroalkyls (10, 24), amide (20), imide (14), carbamate (27), sulfonamide (19), sulfonates (20, 23) and sulfone substituents (28). Substrates with electron-deficient or -neutral aryl groups (14, 19, 22, 28) can be tolerated, but substrates with electron-rich aryl groups result in lower yields (for unsuccessful examples and examples with lower yields, see Table S2).31−35 Some substrates (4, 5, 8, 9) exhibit yields based on recovered starting material (BRSM) greater than 80% but have only moderate yields of isolated pure products due to the incomplete reaction, likely due to product inhibition (see Supporting Information, p S15). Fluorination of unsubstituted quinoline (3) resulted in a 2:1 ratio of C4 and C2 products, slightly favoring the more electrophilic C4 site. A 4.5:1 of C4 and C2 fluorinated products was obtained in the case of 3-acetylquinoline (4), owing to enhanced C4 reactivity caused by electronic effects. The C4 selectivity is inconsistent with the fluorine atom transfer mechanism due to the polarity mismatch.14 The DFT-calculated LUMOs of protonated quinoline derivatives show that C4 has the largest contribution to the LUMO, suggesting that the site selectivity of the fluorination is consistent with the quinoline in the role of the electrophile (Supporting Information, p S73). In general, the selectivity for the C4 position is moderate but can be increased if Lewis acids are used instead of protic acid, presumably to coordinate to the quinoline nitrogen, and thereby sterically disfavor the reaction at C2 (see Supporting Information, pp S22–S23). Benzoquinolines such as 6 and 21 can be fluorinated successfully. Pyridine derivatives are not suitable substrates for the energy transfer because their ET values (79.0 kcal/mol for pyridine)40 are too high for common photosensitizers such as xanthone (ET = 73.8 kcal/mol).40 However, when direct excitation conditions were applied to access triplet-state pyridine–HCl complexes, several pyridine derivatives (25, 26) could be fluorinated successfully.
Figure 4.
Substrate scope. aAzaarene (0.4 mmol), HCl (1.3 equiv); then Selectfluor (3–4 equiv), xanthone (20 mol %), Et3N·HCl (1 equiv), MeCN (0.1 M), 365 nm LED (12 W), 30 °C, 24 h. bYield BRSM is more than 80%. cYields were determined by NMR. d0.1 mmol scale. e1 mmol scale. f280 nm LED (10 W) and a quartz reaction vessel. All isomers of products or their hydrochloride salts have been isolated and characterized as analytically pure samples.
The strategy of electron-transfer-enabled concerted nucleophilic aromatic substitution provides a conceptually new approach to azaarene C–H fluorination. Our work demonstrates how ion-coupled electron transfer avoids the formation of high-energy Meisenheimer intermediates and, therefore, can provide a hypothesis for the design of efficient catalysis by coupling ion transfer with redox processes, which may serve as a novel mechanistic basis for other nucleophilic aromatic C–H functionalization.
Acknowledgments
We thank J. Mateos, J. Kim, S. Lin, B. A. van der Worp, C. Li, and R. Halder (MPI KOFO) for helpful discussions. We thank G. Breitenbruch and N. Sauerborn for the help on preparative HPLC separations; H. Hinrichs, S. Klimek, P. Münstermann, and R. Petzold for HPLC analysis; N. Nöthling for X-ray analysis; M. Leutzsch for NMR spectroscopy analysis; and L. Torkowski for the synthesis of starting material 10 (all from MPI KOFO). L.Z. acknowledges the Alexander von Humboldt Foundation for a Humboldt Research Fellowship. We thank the MPI für Kohlenforschung for funding.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.3c07119.
Experimental procedures and spectral data (PDF)
Open access funded by Max Planck Society.
The authors declare no competing financial interest.
Supplementary Material
References
- Hiyama T.; Yamamoto H.. Organofluorine Building Blocks. In Organofluorine Compounds; Springer: Berlin, 2000; pp 77–118. [Google Scholar]
- de la Torre B. G.; Albericio F. The Pharmaceutical Industry in 2021. An Analysis of FDA Drug Approvals from the Perspective of Molecules. Molecules 2022, 27, 1075. 10.3390/molecules27031075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li J.; Zhang X.; Jin H.; Fan J.; Flores H.; Perlmutter J. S.; Tu Z. Synthesis of fluorine-containing phosphodiesterase 10A (PDE10A) inhibitors and the in vivo evaluation of F-18 labeled PDE10A PET tracers in rodent and nonhuman primate. J. Med. Chem. 2015, 58, 8584–8600. 10.1021/acs.jmedchem.5b01205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shannon R. D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 1976, 32, 751–767. 10.1107/S0567739476001551. [DOI] [Google Scholar]
- Szpera R.; Moseley D. F.; Smith L. B.; Sterling A. J.; Gouverneur V. The Fluorination of C-H bonds: developments and perspectives. Angew. Chem., Int. Ed. 2019, 58, 14824–14848. 10.1002/anie.201814457. [DOI] [PubMed] [Google Scholar]
- Zhang L.; Ritter T. A perspective on late-stage aromatic C-H bond functionalization. J. Am. Chem. Soc. 2022, 144, 2399–2414. 10.1021/jacs.1c10783. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamamoto K.; Li J.; Garber J. A.; Rolfes J. D.; Boursalian G. B.; Borghs J. C.; Genicot C.; Jacq J.; van Gastel M.; Neese F.; Ritter T. Palladium-catalysed electrophilic aromatic C-H fluorination. Nature 2018, 554, 511–514. 10.1038/nature25749. [DOI] [PubMed] [Google Scholar]
- Li J.; Chen J.; Sang R.; Ham W. S.; Plutschack M. B.; Berger F.; Chabbra S.; Schnegg A.; Genicot C.; Ritter T. Photoredox catalysis with aryl sulfonium salts enables site-selective late-stage fluorination. Nat. Chem. 2020, 12, 56–62. 10.1038/s41557-019-0353-3. [DOI] [PubMed] [Google Scholar]
- Chen W.; Huang Z.; Tay N. E.; Giglio B.; Wang M.; Wang H.; Wu Z.; Nicewicz D. A.; Li Z. Direct arene C-H fluorination with 18F- via organic photoredox catalysis. Science 2019, 364, 1170–1174. 10.1126/science.aav7019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fier P. S.; Hartwig J. F. Selective C-H fluorination of pyridines and diazines inspired by a classic amination reaction. Science 2013, 342, 956–960. 10.1126/science.1243759. [DOI] [PubMed] [Google Scholar]
- Fier P. S.; Hartwig J. F. Synthesis and late-stage functionalization of complex molecules through C-H fluorination and nucleophilic aromatic substitution. J. Am. Chem. Soc. 2014, 136, 10139–10147. 10.1021/ja5049303. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Murakami K.; Yamada S.; Kaneda T.; Itami K. C-H functionalization of azines. Chem. Rev. 2017, 117, 9302–9332. 10.1021/acs.chemrev.7b00021. [DOI] [PubMed] [Google Scholar]
- Zhou F. Y.; Jiao L. Recent developments in transition-metal-free functionalization and derivatization reactions of pyridines. Synlett 2021, 32, 159–178. 10.1055/s-0040-1706552. [DOI] [Google Scholar]
- Proctor R. S.; Phipps R. J. Recent advances in Minisci-type reactions. Angew. Chem., Int. Ed. 2019, 58, 13666–13699. 10.1002/anie.201900977. [DOI] [PubMed] [Google Scholar]
- Holmberg-Douglas N.; Nicewicz D. A. Photoredox-catalyzed C-H functionalization reactions. Chem. Rev. 2022, 122, 1925–2016. 10.1021/acs.chemrev.1c00311. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fujiwara Y.; Dixon J. A.; O’Hara F.; Funder E. D.; Dixon D. D.; Rodriguez R. A.; Baxter R. D.; Herlé B.; Sach N.; Collins M. R.; Ishihara Y.; Baran P. S. Practical and innate carbon-hydrogen functionalization of heterocycles. Nature 2012, 492, 95–99. 10.1038/nature11680. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim J. H.; Constantin T.; Simonetti M.; Llaveria J.; Sheikh N. S.; Leonori D. A radical approach for the selective C-H borylation of azines. Nature 2021, 595, 677–683. 10.1038/s41586-021-03637-6. [DOI] [PubMed] [Google Scholar]
- Liu S.; Pan P.; Fan H.; Li H.; Wang W.; Zhang Y. Photocatalytic C-H silylation of heteroarenes by using trialkylhydrosilanes. Chem. Sci. 2019, 10, 3817–3825. 10.1039/C9SC00046A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Josephitis C. M.; Nguyen H. M.; McNally A. Late-Stage C-H Functionalization of Azines. Chem. Rev. 2023, 123, 7655–7691. 10.1021/acs.chemrev.2c00881. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang X.; Nottingham K. G.; Patel C.; Alegre-Requena J. V.; Levy J. N.; Paton R. S.; McNally A. Phosphorus-mediated sp2-sp3 couplings for C-H fluoroalkylation of azines. Nature 2021, 594, 217–222. 10.1038/s41586-021-03567-3. [DOI] [PubMed] [Google Scholar]
- Hilton M. C.; Zhang X.; Boyle B. T.; Alegre-Requena J. V.; Paton R. S.; McNally A. Heterobiaryl synthesis by contractive C-C coupling via P(V) intermediates. Science 2018, 362, 799–804. 10.1126/science.aas8961. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Campbell M. G.; Ritter T. Modern carbon-fluorine bond forming reactions for aryl fluoride synthesis. Chem. Rev. 2015, 115, 612–633. 10.1021/cr500366b. [DOI] [PubMed] [Google Scholar]
- Watson D. A.; Su M.; Teverovskiy G.; Zhang Y.; García-Fortanet J.; Kinzel T.; Buchwald S. L. Formation of ArF from LPdAr (F): catalytic conversion of aryl triflates to aryl fluorides. Science 2009, 325, 1661–1664. 10.1126/science.1178239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Neumann C. N.; Hooker J. M.; Ritter T. Concerted nucleophilic aromatic substitution with 19F- and 18F-. Nature 2016, 534, 369–373. 10.1038/nature17667. [DOI] [PubMed] [Google Scholar]
- Ghiazza C.; Faber T.; Gómez-Palomino A.; Cornella J. Deaminative chlorination of aminoheterocycles. Nat. Chem. 2022, 14, 78–84. 10.1038/s41557-021-00812-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- See Y. Y.; Morales-Colón M. T.; Bland D. C.; Sanford M. S. Development of SNAr nucleophilic fluorination: a fruitful academia-industry collaboration. Acc. Chem. Res. 2020, 53, 2372–2383. 10.1021/acs.accounts.0c00471. [DOI] [PubMed] [Google Scholar]
- Kwan E. E.; Zeng Y.; Besser H. A.; Jacobsen E. N. Concerted nucleophilic aromatic substitutions. Nat. Chem. 2018, 10, 917–923. 10.1038/s41557-018-0079-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rohrbach S.; Smith A. J.; Pang J. H.; Poole D. L.; Tuttle T.; Chiba S.; Murphy J. A. Concerted nucleophilic aromatic substitution reactions. Angew. Chem., Int. Ed. 2019, 58, 16368–16388. 10.1002/anie.201902216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tyburski R.; Liu T.; Glover S. D.; Hammarström L. Proton-coupled electron transfer guidelines, fair and square. J. Am. Chem. Soc. 2021, 143, 560–576. 10.1021/jacs.0c09106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Weinberg D. R.; Gagliardi C. J.; Hull J. F.; Murphy C. F.; Kent C. A.; Westlake B. C.; Paul A.; Ess D. H.; McCafferty D. G.; Meyer T. J. Proton-coupled electron transfer. Chem. Rev. 2012, 112, 4016–4093. 10.1021/cr200177j. [DOI] [PubMed] [Google Scholar]
- Boursalian G. B.; Ham W. S.; Mazzotti A. R.; Ritter T. Charge-transfer-directed radical substitution enables para-selective C-H functionalization. Nat. Chem. 2016, 8, 810–815. 10.1038/nchem.2529. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Serpier F.; Pan F.; Ham W. S.; Jacq J.; Genicot C.; Ritter T. Selective Methylation of Arenes: A Radical C-H Functionalization/Cross-Coupling Sequence. Angew. Chem., Int. Ed. 2018, 57, 10697–10701. 10.1002/anie.201804628. [DOI] [PubMed] [Google Scholar]
- Aguilar Troyano F. J.; Merkens K.; Gómez-Suárez A. Selectfluor® Radical Dication (TEDA2+.)-A Versatile Species in Modern Synthetic Organic Chemistry. Asian J. Org. Chem. 2020, 9, 992–1007. 10.1002/ajoc.202000196. [DOI] [Google Scholar]
- Xiang M.; Xin Z. K.; Chen B.; Tung C. H.; Wu L. Z. Exploring the reducing ability of organic dye (Acr+-Mes) for fluorination and oxidation of benzylic C (sp3)-H bonds under visible light irradiation. Org. Lett. 2017, 19, 3009–3012. 10.1021/acs.orglett.7b01270. [DOI] [PubMed] [Google Scholar]
- Ventre S.; Petronijevic F. R.; MacMillan D. W. Decarboxylative fluorination of aliphatic carboxylic acids via photoredox catalysis. J. Am. Chem. Soc. 2015, 137, 5654–5657. 10.1021/jacs.5b02244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rössler S. L.; Jelier B. J.; Magnier E.; Dagousset G.; Carreira E. M.; Togni A. Pyridinium Salts as Redox-Active Functional Group Transfer Reagents. Angew. Chem., Int. Ed. 2020, 59, 9264–9280. 10.1002/anie.201911660. [DOI] [PubMed] [Google Scholar]
- Wang G.; Zhang H.; Zhao J.; Li W.; Cao J.; Zhu C.; Li S. Homolytic cleavage of a B-B bond by the cooperative catalysis of two Lewis bases: computational design and experimental verification. Angew. Chem., Int. Ed. 2016, 55, 5985–5989. 10.1002/anie.201511917. [DOI] [PubMed] [Google Scholar]
- Kim J.; Sun X.; van der Worp B. A.; Ritter T. Anti-Markovnikov hydrochlorination and hydronitrooxylation of α-olefins. Nat. Catal. 2023, 6, 196–203. 10.1038/s41929-023-00914-7. [DOI] [Google Scholar]
- Strieth-Kalthoff F.; James M. J.; Teders M.; Pitzer L.; Glorius F. Energy transfer catalysis mediated by visible light: principles, applications, directions. Chem. Soc. Rev. 2018, 47, 7190–7202. 10.1039/C8CS00054A. [DOI] [PubMed] [Google Scholar]
- Morov S. L.; Carmicheal I.; Hug G. L.. Handbook of Photochemistry; Marcel Dekker, Inc.: New York, 1993. [Google Scholar]
- Ma J.; Chen S.; Bellotti P.; Guo R.; Schäfer F.; Heusler A.; Zhang X.; Daniliuc C.; Brown M. K.; Houk K. N.; Glorius F. Photochemical intermolecular dearomative cycloaddition of bicyclic azaarenes with alkenes. Science 2021, 371, 1338–1345. 10.1126/science.abg0720. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Neese F. The ORCA program system. WIREs Comput. Mol. Sci. 2012, 2, 73–78. 10.1002/wcms.81. [DOI] [Google Scholar]
- Pitts C. R.; Ling B.; Woltornist R.; Liu R.; Lectka T. Triethylborane-initiated radical chain fluorination: a synthetic method derived from mechanistic insight. J. Org. Chem. 2014, 79, 8895–8899. 10.1021/jo501520e. [DOI] [PubMed] [Google Scholar]
- Xia J. B.; Zhu C.; Chen C. Visible light-promoted metal-free C-H activation: diarylketone-catalyzed selective benzylic mono-and difluorination. J. Am. Chem. Soc. 2013, 135, 17494–17500. 10.1021/ja410815u. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Markle T. F.; Darcy J. W.; Mayer J. M. A new strategy to efficiently cleave and form C-H bonds using proton-coupled electron transfer. Sci. Adv. 2018, 4, eaat5776 10.1126/sciadv.aat5776. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhu X. Q.; Yang Y.; Zhang M.; Cheng J. P. First estimation of C4-H bond dissociation energies of NADH and its radical cation in aqueous solution. J. Am. Chem. Soc. 2003, 125, 15298–15299. 10.1021/ja0385179. [DOI] [PubMed] [Google Scholar]
- Zhu X. Q.; Li H. R.; Li Q.; Ai T.; Lu J. Y.; Yang Y.; Cheng J. P. Determination of the C4-H Bond Dissociation Energies of NADH Models and Their Radical Cations in Acetonitrile. Chem. Eur. J. 2003, 9, 871–880. 10.1002/chem.200390108. [DOI] [PubMed] [Google Scholar]
- Labet V.; Morell C.; Toro-Labbé A.; Grand A. Is an elementary reaction step really elementary? Theoretical decomposition of asynchronous concerted mechanisms. Phys. Chem. Chem. Phys. 2010, 12, 4142–4151. 10.1039/b924589h. [DOI] [PubMed] [Google Scholar]
- Poulet G.; Laverdet G.; Jourdain J. L.; Le Bras G. Kinetic study of the reactions of acetonitrile with chlorine (Cl) and hydroxyl radicals. J. Phys. Chem. 1984, 88, 6259–6263. 10.1021/j150669a041. [DOI] [Google Scholar]
- Komura A.; Uchida K.; Yagi M.; Higuchi J. Electron spin resonance and phosphorescence of quinoline, isoquinoline and their protonated cations in the phosphorescent triplet states. J. Photochem. Photobiol., A 1988, 42, 293–300. 10.1016/1010-6030(88)80072-8. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.