Radical fluorination

Last updated

Radical fluorination is a type of fluorination reaction, complementary to nucleophilic and electrophilic approaches. [1] It involves the reaction of an independently generated carbon-centered radical with an atomic fluorine source and yields an organofluorine compound.

Contents

Radical fluorination.tif

Historically, only three atomic fluorine sources were available for radical fluorination: Fluorine (F2), hypofluorites (O–F based reagents) and XeF2. Their high reactivity, and the difficult handling of F2 and the hypofluorites, limited the development of radical fluorination compared to electrophilic and nucleophilic methods. [2] The uncovering of the ability of electrophilic N–F fluorinating agents to act as an atomic fluorine source [3] led to a renaissance in radical fluorination. [2]

Various methodologies have since been developed for the radical formation of C–F bonds. [1] The radical intermediates have been generated from carboxylic acids and boronic acid derivatives, by radical addition to alkenes, or C–H and C–C bond activations. New sources of atomic fluorine are now emerging, such as metal fluoride complexes.

Sources of atomic fluorine

Fluorine gas

Fluorine gas (F2) can act both as an electrophilic and atomic source of fluorine. [4] The weak F–F bond strength (36 kcal/mol (150 kJ/mol) [5] ) allows for homolytic cleavage. The reaction of F2 with organic compounds is, however, highly exothermic and can lead to non-selective fluorinations and C–C cleavage, as well as explosions. [6] Only a few selective radical fluorination methods have been reported. [7] [8] The use of fluorine for radical fluorination is mainly limited to perfluorination reactions. [5]

O–F reagents

The O–F bond of hypofluorites is relatively weak. For trifluoromethyl hypofluorite (CF3OF), it has been estimated to be 43.5 kcal/mol (182 kJ/mol). [9] The ability of trifluoromethyl hypofluorite to transfer fluorine to alkyl radicals is notably demonstrated by reacting independently generated ethyl radicals from ethene and tritium in the presence of CF3OF. [10] The high reactivity of hypofluorites has limited their application to selective radical fluorination. They can, however, be used as radical initiators for polymerization. [11]

XeF2

Xenon difluoride (XeF2) has mainly been used for radical fluorination in radical decarboxylative fluorination reactions. [12] In this Hunsdiecker-type reaction, xenon difluoride is used to generate the radical intermediate, as well as the fluorine transfer source. [13]

Xe decarboxylation.tif

XeF2 can also be used to generate aryl radicals from arylsilanes, and act as an atomic fluorine source to furnish aryl fluorides. [14]

Xe silanes.tif

N–F reagents

Selectfluor and N-fluorobenzenesulfonimide (NFSI) are traditionally used as electrophilic sources of fluorine, but their ability to transfer fluorine to alkyl radicals has recently been demonstrated. [3] They are now commonly used as fluorine transfer agents to alkyl radicals. [1]

Others

Examples of radical fluorination using bromine trifluoride (BrF3) [15] and fluorinated solvents [16] have been reported. Recent examples in radical fluorination suggest that in-situ generated metal fluoride complexes can also act as fluorine transfer agents to alkyl radicals.[ citation needed ]

Radical fluorination methodologies

Decarboxylative fluorination

The thermolysis of t-butyl peresters has been used to generate alkyl radicals in presence of NFSI and Selectfluor. [3] The radicals' intermediates were efficiently fluorinated, demonstrating the ability of the two electrophilic fluorinating agents to transfer fluorine to alkyl radicals.

Perester.tif

Carboxylic acids can be used as radical precursors in radical fluorination methods. Metal catalysts such as silver [17] and manganese [18] have been used to induce the fluorodecarboxylation. The fluorodecarboxylation of carboxylic acids can also be triggered using photoredox catalysis. [19] [20] More specifically, phenoxyacetic acid derivatives have been shown to undergo fluorodecarboxylation when directly exposed to ultraviolet irradiation [21] or via the use of a photosensitizer. [22]

Carboxylic acid.tif

Radical fluorination of alkenes

Alkyl radicals generated from radical additions to alkenes have also been fluorinated. Hydrides [23] and nitrogen-, [24] carbon-, [25] and phosphorus-centered [26] radicals have been employed, yielding a wide range of fluorinated difunctionalized compounds.

Fluorination of boronic acid derivatives

Alkyl fluorides have been synthesized via radicals generated from boronic acid derivatives using silver. [27]

Radical fluorination of boronates.tif

C(sp3)–H fluorination

One major advantage of radical fluorination is that it allows the direct fluorination of remote C–H bonds. Metal catalysts such as manganese, [28] copper, [29] and tungsten [30] have been used to promote the reaction. Metal-free C(sp3)–H fluorinations rely on the use of radical initiators (triethylborane, [31] persulfates [32] or N-oxyl radicals [33] ) or organic photocatalysts. [33]

Some methods have also been developed to selectively fluorinate benzylic C–H bonds. [34]

C–C bond activation

Cyclobutanols and cyclopropanols have been used as radical precursors for the synthesis of β- or γ-fluoroketones. The strained rings undergo C–C bond cleavage in presence of a silver [35] [36] or an iron catalyst [36] or when exposed to ultraviolet light in presence of a photosensitizer. [37]

C--C bond activation.tif

Potential applications

One potential application of radical fluorination is for efficiently accessing novel moieties to serve as building blocks in medicinal chemistry. [38] Derivatives of propellane with reactive functional groups, such as the hydrochloride salt of 3-fluorobicyclo[1.1.1]pentan-1-amine, are accessible by this approach. [38]

Related Research Articles

Haloalkane Group of chemical compounds derived from alkanes containing one or more halogens

The haloalkanes known as halogenoalkanes' or alkyl halides) are alkanes containing one or more halogen substituents. They are a subset of the general class of halocarbons, although the distinction is not often made. Haloalkanes are widely used commercially. They are used as flame retardants, fire extinguishants, refrigerants, propellants, solvents, and pharmaceuticals. Subsequent to the widespread use in commerce, many halocarbons have also been shown to be serious pollutants and toxins. For example, the chlorofluorocarbons have been shown to lead to ozone depletion. Methyl bromide is a controversial fumigant. Only haloalkanes that contain chlorine, bromine, and iodine are a threat to the ozone layer, but fluorinated volatile haloalkanes in theory may have activity as greenhouse gases. Methyl iodide, a naturally occurring substance, however, does not have ozone-depleting properties and the United States Environmental Protection Agency has designated the compound a non-ozone layer depleter. For more information, see Halomethane. Haloalkane or alkyl halides are the compounds which have the general formula "RX" where R is an alkyl or substituted alkyl group and X is a halogen.

In chemistry, halogenation is a chemical reaction that entails the introduction of one or more halogens into a compound. Halide-containing compounds are pervasive, making this type of transformation important, e.g. in the production of polymers, drugs. This kind of conversion is in fact so common that a comprehensive overview is challenging. This article mainly deals with halogenation using elemental halogens (F2, Cl2, Br2, I2). Halides are also commonly introduced using salts of the halides and halogen acids. Many specialized reagents exist for and introducing halogens into diverse substrates, e.g. thionyl chloride.

The Sandmeyer reaction is a chemical reaction used to synthesize aryl halides from aryl diazonium salts using copper salts as reagents or catalysts. It is an example of a radical-nucleophilic aromatic substitution. The Sandmeyer reaction provides a method through which one can perform unique transformations on benzene, such as halogenation, cyanation, trifluoromethylation, and hydroxylation.

Organoboron chemistry

Organoborane or organoboron compounds are chemical compounds of boron and carbon that are organic derivatives of BH3, for example trialkyl boranes. Organoboron chemistry or organoborane chemistry is the chemistry of these compounds.

<i>N</i>,<i>N</i>-Diisopropylethylamine chemical compound

N,N-Diisopropylethylamine, or Hünig's base, is an organic compound and an amine. It is named after the German chemist Siegfried Hünig. It is used in organic chemistry as a base. It is commonly abbreviated as DIPEA,DIEA, or i-Pr2NEt.

The Hunsdiecker reaction is a name reaction in organic chemistry whereby silver salts of carboxylic acids react with a halogen to produce an organic halide. It is an example of both a decarboxylation and a halogenation reaction as the product has one fewer carbon atoms than the starting material and a halogen atom is introduced its place. The reaction was first demonstrated by Alexander Borodin in his 1861 reports of the preparation of methyl bromide from silver acetate. Shortly after, the approach was applied to the degradation of fatty acids in the laboratory of Adolf Lieben. However, it is named for Cläre Hunsdiecker and her husband Heinz Hunsdiecker, whose work in the 1930s developed it into a general method. Several reviews have been published, and a catalytic approach has been developed.

Xenon difluoride Chemical compound

Xenon difluoride is a powerful fluorinating agent with the chemical formula XeF
2
, and one of the most stable xenon compounds. Like most covalent inorganic fluorides it is moisture-sensitive. It decomposes on contact with water vapor, but is otherwise stable in storage. Xenon difluoride is a dense, colourless crystalline solid.

Organozinc compound

Organozinc compounds in organic chemistry contain carbon to zinc chemical bonds. Organozinc chemistry is the science of organozinc compounds describing their physical properties, synthesis and reactions.

Organofluorine chemistry describes the chemistry of the organofluorines, organic compounds that contain the carbon–fluorine bond. Organofluorine compounds find diverse applications ranging from oil and water repellents to pharmaceuticals, refrigerants, and reagents in catalysis. In addition to these applications, some organofluorine compounds are pollutants because of their contributions to ozone depletion, global warming, bioaccumulation, and toxicity. The area of organofluorine chemistry often requires special techniques associated with the handling of fluorinating agents.

The Kulinkovich reaction describes the organic synthesis of cyclopropanols via reaction of esters with dialkyldialkoxytitanium reagents, generated in situ from Grignard reagents bearing hydrogen in beta-position and titanium(IV) alkoxides such as titanium isopropoxide. This reaction was first reported by Oleg Kulinkovich and coworkers in 1989.

Cyanuric fluoride Chemical compound

Cyanuric fluoride or 2,4,6-trifluoro-1,3,5-triazine is a chemical compound with the formula (CNF)3. It is a colourless, pungent liquid. It has been used as a precursor for fibre-reactive dyes, as a specific reagent for tyrosine residues in enzymes, and as a fluorinating agent.

1-Chloromethyl-4-fluoro-1,4-diazoniabicyclo[2.2.2]octane bis(tetrafluoroborate) or Selectfluor, a trademark of Air Products and Chemicals, is a reagent in chemistry that is used as a fluorine donor. This compound is a derivative of the nucleophillic base DABCO. This colourless salt was first described in 1992 and has since been commercialized for use in organofluorine chemistry for electrophilic fluorination.

Electrochemical fluorination (ECF), or electrofluorination, is a foundational organofluorine chemistry method for the preparation of fluorocarbon-based organofluorine compounds. The general approach represents an application of electrosynthesis. The fluorinated chemical compounds produced by ECF are useful because of their distinctive solvation properties and the relative inertness of carbon–fluorine bonds. Two ECF synthesis routes are commercialized and commonly applied: the Simons process and the Phillips Petroleum process. It is also possible to electrofluorinate in various organic media. Prior to the development of these methods, fluorination with fluorine, a dangerous oxidant, was a dangerous and wasteful process. Also, ECF can be cost-effective but it may also result in low yields.

Organogold chemistry is the study of compounds containing gold–carbon bonds. They are studied in academic research, but have not received widespread use otherwise. The dominant oxidation states for organogold compounds are I with coordination number 2 and a linear molecular geometry and III with CN = 4 and a square planar molecular geometry. The first organogold compound discovered was gold(I) carbide Au2C2, which was first prepared in 1900.

Thionyl tetrafluoride Chemical compound

Thionyl tetrafluoride is an inorganic compound gas with the formula SOF4. It is also known as sulfur tetrafluoride oxide. The shape of the molecule is a distorted trigonal bipyramid, with the oxygen found on the equator. The atoms on the equator have shorter bond lengths than the fluorine atoms on the axis. The sulfur oxygen bond is 1.409Å. A S−F bond on the axis has length 1.596Å and the S−F bond on the equator has length 1.539Å. The angle between the equatorial fluorine atoms is 112.8°. The angle between axial fluorine and oxygen is 97.7°. The angle between oxygen and equatorial fluorine is 123.6° and between axial and equatorial fluorine is 85.7°. The fluorine atoms only produce one NMR line, probably because they exchange positions.

Metal carbon dioxide complexes are coordination complexes that contain carbon dioxide ligands. Aside from the fundamental interest in the coordination chemistry of simple molecules, studies in this field are motivated by the possibility that transition metals might catalyze useful transformations of CO2. This research is relevant both to organic synthesis and to the production of "solar fuels" that would avoid the use of petroleum-based fuels.

Trifluoromethylation in organic chemistry describes any organic reaction that introduces a trifluoromethyl group in an organic compound. Trifluoromethylated compounds are of some importance in pharmaceutical industry and agrochemicals. Several notable pharmaceutical compounds have a trifluoromethyl group incorporated: fluoxetine, mefloquine, Leflunomide, nulitamide, dutasteride, bicalutamide, aprepitant, celecoxib, fipronil, fluazinam, penthiopyrad, picoxystrobin, fluridone, norflurazon, sorafenib and triflurazin. A relevant agrochemical is trifluralin. The development of synthetic methods for adding trifluoromethyl groups to chemical compounds is actively pursued in academic research.

Catellani reaction

The Catellani reaction was discovered by Marta Catellani and co-workers in 1997. The reaction uses aryl iodides to perform bi- or tri-functionalization, including C-H functionalization of the unsubstituted ortho position(s), followed a terminating cross-coupling reaction at the ipso position. This cross-coupling cascade reaction depends on the ortho-directing transient mediator, norbornene.

The Mukaiyama hydration is an organic reaction involving formal addition of an equivalent of water across an olefin by the action of catalytic bis(acetylacetonato)cobalt(II) complex, phenylsilane and atmospheric oxygen to produce an alcohol with Markovnikov selectivity.

An organic azide is organic compounds containing the azide (N3) functional group. Because of the hazards associated with their use, few azides are used commercially although they exhibit interesting reactivity for researchers. Low molecular weight azides are considered especially hazardous and are avoided. In the research laboratory, azides are precursors to amines. They are also popular for their participation in the "click reaction" and in Staudinger ligation. These two reactions are generally quite reliable, lending themselves to combinatorial chemistry.

References

  1. 1 2 3 Paquin, Jean-François; Sammis, Glenn; Chatalova-Sazepin, Claire; Hemelaere, Rémy (Aug 2015). "Recent Advances in Radical Fluorination". Synthesis. 47 (17): 2554–2569. doi:10.1055/s-0034-1378824.
  2. 1 2 Sibi, Mukund P.; Landais, Yannick (Feb 2013). "Csp3–F Bond Formation: A Free-Radical Approach". Angewandte Chemie International Edition. 52 (13): 3570–3572. doi:10.1002/anie.201209583. PMID   23441011.
  3. 1 2 3 Rueda-Becerril, Montserrat; Chatalova-Sazepin, Claire; Leung, Joe C. T.; Okbinoglu, Tulin; Kennepohl, Pierre; Paquin, Jean-François; Sammis, Glenn M. (Mar 2012). "Fluorine Transfer to Alkyl Radicals". Journal of the American Chemical Society. 134 (9): 4026–4029. doi:10.1021/ja211679v. ISSN   0002-7863. PMID   22320293.
  4. Bigelow, Lucius A. (Feb 1947). "The Action of Elementary Fluorine upon Organic Compounds". Chemical Reviews. 40 (1): 51–115. doi:10.1021/cr60125a004. ISSN   0009-2665. PMID   20287884.
  5. 1 2 Hutchinson, John; Sandford, Graham (1997). Chambers, Richard D. (ed.). Elemental Fluorine in Organic Chemistry. Topics in Current Chemistry. Berlin, Heidelberg: Springer. pp. 1–43. doi:10.1007/3-540-69197-9_1. ISBN   978-3-540-63170-5.
  6. Simons, J. H.; Block, L. P. (Oct 1939). "Fluorocarbons. The Reaction of Fluorine with Carbon". Journal of the American Chemical Society. 61 (10): 2962–2966. doi:10.1021/ja01265a111. ISSN   0002-7863.
  7. Grakauskas, Vytautas (Aug 1969). "Aqueous fluorination of carboxylic acid salts". The Journal of Organic Chemistry. 34 (8): 2446–2450. doi:10.1021/jo01260a040. ISSN   0022-3263.
  8. Bockemüller, Wilhelm (Jan 1933). "Versuche zur Fluorierung organischer Verbindungen. III. Über die Einwirkung von Fluor auf organische Verbindungen" [Attempts at fluorination of organic compounds. III. On the effect of fluorine upon organic compounds]. Justus Liebigs Annalen der Chemie (in German). 506 (1): 20–59. doi:10.1002/jlac.19335060103. ISSN   1099-0690.
  9. Czarnowski, J.; Castellano, E.; Schumacher, H. J. (Jan 1968). "The energy of the O–F bond in trifluoromethyl hypofluorite". Chemical Communications (20): 1255. doi:10.1039/c19680001255.
  10. Wang, Nunyii; Rowland, F. S. (Nov 1985). "Trifluoromethyl hypofluorite: a fluorine-donating radical scavenger". The Journal of Physical Chemistry. 89 (24): 5154–5155. doi:10.1021/j100270a006. ISSN   0022-3654.
  11. Francesco, Venturini; Sansotera, Maurizio; Navarrini, Walter (November 2013). "Recent developments in the chemistry of organic perfluoro hypofluorites". Journal of Fluorine Chemistry. 2013 ACS Fluorine Award Issue: Professor Iwao Ojima. 155: 2–20. doi:10.1016/j.jfluchem.2013.07.005.
  12. Tius, Marcus A. (Jun 1995). "Xenon difluoride in synthesis". Tetrahedron. 51 (24): 6605–6634. doi:10.1016/0040-4020(95)00362-C.
  13. Patrick, Timothy B.; Darling, Diana L. (Aug 1986). "Fluorination of activated aromatic systems with cesium fluoroxysulfate". The Journal of Organic Chemistry. 51 (16): 3242–3244. doi:10.1021/jo00366a044. ISSN   0022-3263.
  14. Lothian, Aileen P.; Ramsden, Christopher A. (Jan 1993). "Rapid Fluorodesilylation of Aryltrimethylsilanes Using Xenon Difuoride: An Efficient New Route to Aromatic Fluorides". Synlett. 1993 (10): 753–755. doi:10.1055/s-1993-22596.
  15. Sasson, Revital; Rozen, Shlomo (Jan 2005). "Constructing the CF3 group; unique trifluorodecarboxylation induced by BrF3". Tetrahedron. 61 (5): 1083–1086. doi:10.1016/j.tet.2004.11.063.
  16. Yamada, Shigeyuki; Gavryushin, Andrei; Knochel, Paul (March 2010). "Convenient Electrophilic Fluorination of Functionalized Aryl and Heteroaryl Magnesium Reagents". Angewandte Chemie International Edition. 49 (12): 2215–2218. doi:10.1002/anie.200905052. PMID   20162637.
  17. Yin, Feng; Wang, Zhentao; Li, Zhaodong; Li, Chaozhong (Jun 2012). "Silver-Catalyzed Decarboxylative Fluorination of Aliphatic Carboxylic Acids in Aqueous Solution". Journal of the American Chemical Society. 134 (25): 10401–10404. doi:10.1021/ja3048255. ISSN   0002-7863. PMID   22694301.
  18. Huang, Xiongyi; Liu, Wei; Hooker, Jacob M.; Groves, John T. (Apr 2015). "Targeted Fluorination with the Fluoride Ion by Manganese-Catalyzed Decarboxylation". Angewandte Chemie International Edition. 54 (17): 5241–5245. doi:10.1002/anie.201500399. ISSN   1521-3773. PMID   25736895.
  19. Rueda Becerril, Montserrat; Mahé, Olivier; Drouin, Myriam; Majewski, Marek B.; West, Julian G.; Wolf, Michael O.; Sammis, Glenn M.; Paquin, Jean-François (Jan 2014). "Direct C–F Bond Formation Using Photoredox Catalysis". Journal of the American Chemical Society. 136 (6): 2637–2641. doi:10.1021/ja412083f. PMID   24437369.
  20. Ventre, Sandrine; Petronijević, Filip R.; MacMillan, David W. C. (Apr 2015). "Decarboxylative Fluorination of Aliphatic Carboxylic Acids via Photoredox Catalysis". Journal of the American Chemical Society. 137 (17): 5654–5657. doi:10.1021/jacs.5b02244. PMC   4862610 . PMID   25881929.
  21. Leung, Joe C. T.; Chatalova-Sazepin, Claire; West, Julian G.; Rueda Becerril, Montserrat; Paquin, Jean-François; Sammis, Glenn M. (Oct 2012). "Photo-fluorodecarboxylation of 2-Aryloxy and 2-Aryl Carboxylic Acids". Angewandte Chemie International Edition. 51 (43): 10804–10807. doi:10.1002/anie.201206352. ISSN   1521-3773. PMID   23023887.
  22. Leung, Joe C. T.; Sammis, Glenn M. (Apr 2015). "Radical Decarboxylative Fluorination of Aryloxyacetic Acids Using N-Fluorobenzenesulfonimide and a Photosensitizer". European Journal of Organic Chemistry. 2015 (10): 2197–2204. doi:10.1002/ejoc.201500038. ISSN   1099-0690.
  23. Barker, Timothy J.; Boger, Dale L. (Aug 2012). "Fe(III)/NaBH4-Mediated Free Radical Hydrofluorination of Unactivated Alkenes". Journal of the American Chemical Society. 134 (33): 13588–13591. doi:10.1021/ja3063716. PMC   3425717 . PMID   22860624.
  24. Li, Zhaodong; Zhang, Chengwei; Zhu, Lin; Liu, Chao; Li, Chaozhong (Feb 2014). "Transition-metal-free, room-temperature radical azidofluorination of unactivated alkenes in aqueous solution". Organic Chemistry Frontiers. 1 (1): 100–104. doi:10.1039/c3qo00037k.
  25. Kindt, Stephanie; Heinrich, Markus R. (Nov 2014). "Intermolecular Radical Carbofluorination of Non-activated Alkenes". Chemistry: A European Journal. 20 (47): 15344–15348. doi:10.1002/chem.201405229. ISSN   1521-3765. PMID   25303212.
  26. Zhang, Chengwei; Li, Zhaodong; Zhu, Lin; Yu, Limei; Wang, Zhentao; Li, Chaozhong (Sep 2013). "Silver-Catalyzed Radical Phosphonofluorination of Unactivated Alkenes". Journal of the American Chemical Society. 135 (38): 14082–14085. doi:10.1021/ja408031s. PMID   24025164.
  27. Li, Zhaodong; Wang, Zhentao; Zhu, Lin; Tan, Xinqiang; Li, Chaozhong (Nov 2014). "Silver-Catalyzed Radical Fluorination of Alkylboronates in Aqueous Solution". Journal of the American Chemical Society. 136 (46): 16439–16443. doi:10.1021/ja509548z. PMID   25350556.
  28. Liu, Wei; Huang, Xiongyi; Cheng, Mu-Jeng; Nielsen, Robert J.; Goddard, William A.; Groves, John T. (Sep 2012). "Oxidative Aliphatic C–H Fluorination with Fluoride Ion Catalyzed by a Manganese Porphyrin" (PDF). Science. 337 (6100): 1322–1325. Bibcode:2012Sci...337.1322L. doi:10.1126/science.1222327. ISSN   0036-8075. PMID   22984066. S2CID   90742.
  29. Bloom, Steven; Pitts, Cody Ross; Miller, David Curtin; Haselton, Nathan; Holl, Maxwell Gargiulo; Urheim, Ellen; Lectka, Thomas (Oct 2012). "A Polycomponent Metal-Catalyzed Aliphatic, Allylic, and Benzylic Fluorination". Angewandte Chemie International Edition. 51 (42): 10580–10583. doi:10.1002/anie.201203642. ISSN   1521-3773. PMID   22976771.
  30. Halperin, Shira D.; Fan, Hope; Chang, Stanley; Martin, Rainer E.; Britton, Robert (Mar 2014). "A Convenient Photocatalytic Fluorination of Unactivated C–H Bonds". Angewandte Chemie International Edition. 53 (18): 4690–4693. doi:10.1002/anie.201400420. PMID   24668727.
  31. Pitts, Cody Ross; Ling, Bill; Woltornist, Ryan; Liu, Ran; Lectka, Thomas (Oct 2014). "Triethylborane-Initiated Radical Chain Fluorination: A Synthetic Method Derived from Mechanistic Insight". The Journal of Organic Chemistry. 79 (18): 8895–8899. doi:10.1021/jo501520e. PMID   25137438.
  32. Zhang, Xiaofei; Guo, Shuo; Tang, Pingping (Jun 2015). "Transition-metal free oxidative aliphatic C–H fluorination". Organic Chemistry Frontiers. 2 (7): 806–810. doi:10.1039/c5qo00095e.
  33. 1 2 Amaoka, Yuuki; Nagatomo, Masanori; Inoue, Masayuki (Apr 2013). "Metal-Free Fluorination of C(sp3)–H Bonds Using a Catalytic N-Oxyl Radical". Organic Letters. 15 (9): 2160–2163. doi:10.1021/ol4006757. PMID   23600550.
  34. Koperniku, Ana; Liu, Hongqiang; Hurley, Paul B. (Jan 2016). "Mono- and Difluorination of Benzylic Carbon Atoms". European Journal of Organic Chemistry. 2016 (5): 871–886. doi:10.1002/ejoc.201501329. ISSN   1099-0690.
  35. Ishida, Naoki; Okumura, Shintaro; Nakanishi, Yuuta; Murakami, Masahiro (Jan 2015). "Ring-opening Fluorination of Cyclobutanols and Cyclopropanols Catalyzed by Silver". Chemistry Letters. 44 (6): 821–823. doi:10.1246/cl.150138.
  36. 1 2 Ren, Shichao; Feng, Chao; Loh, Teck-Peng (Apr 2015). "Iron- or silver-catalyzed oxidative fluorination of cyclopropanols for the synthesis of β-fluoroketones". Organic & Biomolecular Chemistry. 13 (18): 5105–5109. doi:10.1039/c5ob00632e. PMID   25866198.
  37. Bloom, Steven; Bume, Desta Doro; Pitts, Cody Ross; Lectka, Thomas (May 2015). "Site-Selective Approach to β-Fluorination: Photocatalyzed Ring Opening of Cyclopropanols". Chemistry: A European Journal. 21 (22): 8060–8063. doi:10.1002/chem.201501081. ISSN   1521-3765. PMID   25877004.
  38. 1 2 Goh, Y. L.; Adsool, V. A. (Dec 2015). "Radical fluorination powered expedient synthesis of 3-fluorobicyclo[1.1.1]pentan-1-amine". Organic & Biomolecular Chemistry. 13 (48): 11597–11601. doi:10.1039/C5OB02066B. PMID   26553141.