Next Article in Journal
Catalytic Degradation of Toluene over MnO2/LaMnO3: Effect of Phase Type of MnO2 on Activity
Previous Article in Journal
Novel Challenges on the Catalytic Synthesis of 5-Hydroxymethylfurfural (HMF) from Real Feedstocks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Catalytic C–F Bond Activation and Transformation of Aromatic Fluorides

1
Provincial Key Laboratory of Polyolefin New Materials, College of Chemistry & Chemical Engineering, The Northeast Petroleum University, Daqing 163318, China
2
Petrochina Petrochemical Research Institute, Petrochina Company Limited, Beijing 102206, China
3
State Key Laboratory of Fine Chemicals, Department of Pharmaceutical Science, School of Chemical Engineering, Dalian University of Technology, Dalian 116024, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(12), 1665; https://doi.org/10.3390/catal12121665
Submission received: 22 November 2022 / Revised: 14 December 2022 / Accepted: 15 December 2022 / Published: 18 December 2022
(This article belongs to the Topic Catalysis: Homogeneous and Heterogeneous)

Abstract

:
The activation and transformation of C–F bonds in fluoro-aromatics is a highly desirable process in organic chemistry. It provides synthetic methods/protocols for the generation of organic compounds possessing single or multiple C–F bonds, and effective catalytic systems for further study of the activation mode of inert chemical bonds. Due to the high polarity of the C–F bond and it having the highest bond energy in organics, C–F activation often faces considerable academic challenges. In this mini-review, the important research achievements in the activation and transformation of aromatic C–F bond, catalyzed by transition metal and metal-free systems, are presented.

1. Introduction

Organic fluoride refers to a class of organic compounds in which one or more of the C–H bonds are replaced by C–F bonds. Due to the high electronegativity of the fluorine atom and the high C–F bond energy, the fluorine-containing organic molecules usually exhibit unique and interesting chemical and biological properties to a great extent [1] and have been widely used in pharmaceuticals, synthetic materials, agricultural chemicals and fine chemicals [2,3], which play indispensable roles in our daily life. The nonmetallic element fluorine atom with the highest electronegativity (Pauling’s Electronegativity 3.98) in the periodic table of elements is indeed a small atom (Pauling’s Van der Waals radius 135 pm)with increasing impact in organic and biochemistries [4,5,6].
The activation of inert chemical bonds is an ineluctable task in general organic synthesis; however, compared with the rich chemistries of activation/transformation of various single bonds with high bonding energies, such as (Eb kJ/mol) C–H (Eb 414) [7,8,9,10,11,12,13,14,15,16,17,18,19], C–C (Eb 347) [20,21,22,23,24,25,26,27,28,29,30], C–N (Eb 308) [31,32,33,34,35,36,37,38,39,40] and C–O (Eb 358) [41,42,43,44,45,46,47,48,49,50,51,52,53], the activation of C–F bond (Eb 485)—which could be built efficiently following typical procedures [4,54,55,56,57,58,59,60,61,62]—is still scarce, due mainly to the abnormal strength of the bond energy and the shielding effect of fluorine atoms. Developing an efficient methodology for the functionalization of C–F bond is highly desirable for the following both academic and industrial reasons [63,64,65]: (1) It will enrich the catalytic toolbox for the activation of inert and polarized chemical bonds; (2) Selective C–F activation of multiple C–F bond containing compounds can provide attractive strategies for synthetic organic chemistry; (3) Regio-selective conversion of fluoride into various functional groups could provide novel molecular structures that are otherwise difficult to create. Therefore, catalytic C–F activations are receiving increased attention and significant progress has been made in the activation and transformation of C–F bonds in alkane fluorides [66,67,68,69,70,71,72,73,74,75,76,77]. This mini-review focuses on recent progress in the activation and functionalization of aromatic C–F bonds catalyzed by transition metal catalysts, as well some metal-free catalytic systems [78,79,80].

2. Transition Metal-Promoted Catalytic C–F Activation of Aromatic Hydrocarbons

Transition metals catalyst-promoted C–F bond activation and functionalization with high selectivity and activities represent effective ways to convert easily available and low-cost fluoro-aromatics into high value fluorine-containing or fluorine-free products via C–C (carbon–carbon coupling), C–H (hydro-defluorination, HDF) and carbon-hetero-atom bond formation processes [81]. Progress in recent years in aromatic C–F bond activation has accumulated steadily, and the newly developed methods could potentially also be applied to fluoro-containing organic waste degradation.

2.1. Ni-Catalyzed Activation of C–F Bonds in Fluoro-Aromatics

Nickle complexes are widely used in catalytic chemical transformations. The coupling reaction between the Grignard reagent and fluoro-aromatics catalyzed by Nickel (II) complexes containing bidentate phosphine ligands (1,2-diphenylphosphinoethane, DPPE, and 1,3-diphenyl phosphinopropane, DPPP) proved to be an efficient method for the defluorocoupling reaction. Thus, in the presence of a DPPE-Ni(II) complex, Grignard reagent reacts with N-heterocyclic aryl fluorides to produce N-heterocyclic biphenyl compounds (TOF ca. 1.1 h−1,Scheme 1), whereas in the presence of DPPP-NiCl2 under mild conditions, the reaction proceeded smoothly to generate the corresponding unsymmetrical biphenyl compounds with high yield and selectivity (TOF ca. 76 h−1, Scheme 1) [82,83]. The structure of the bidentated phosphino-ligand plays a key role [84].
Cao et al. reported a Na-assisted NiCl2-dppp catalytic coupling reaction between fluoro-aromatics and alkyl zinc reagents. The pyridinyl ring was used as a guiding group to introduce a new C–C bond at the ortho position. The mechanistic study showed that there were free radical intermediates generated during the reaction [85]. The C–C coupling product can be obtained in high yield (TOF ca. 1.2–1.5 h−1, Scheme 2).
Zhang et al. reported the ortho-selective hydrodefluorination (HDF) of the similar fluoro-aromatics. In the presence of silane, using easily available inorganic Ni(II), NiCl2·6H2O, o-2-pyridinyl-fluorobenzene derivatives underwent selective partial HDF reaction (TOF ca. 1.3–2.0 h−1) [86]; this reaction can be applied to a variety of multifluoro-aromatic substrates, and a range of usually hard-to-access partial fluoro-aromatics can be effectively prepared (Scheme 3).
Shibata et al. developed an effective nickel(0)-based catalytic metathesis system for the defluoro-silylation reaction between fluoro-aromatics and R3SiBpin, i.e., an aromatic C–F bond can be activated in the absence of additional ligands to obtain aryl silanes in high yield. This reaction is suitable for various inert fluoro-aromatics and a wide range of substrates [87]. Interestingly, it was further observed that the reaction can proceed under very similar conditions in the absence of a Ni(0) catalyst (Scheme 4).
In 2018, Sawamura et al. reported the cross-coupling reaction between fluoro-aromatics and primary amines, mediated by Ni-complexes and producing secondary amines in high selectivity [88]. Electron-rich aryl fluoride 4-fluoroanisole reacted with alkyl-amine to give secondary amine products in up to 94% isolated yields (TOF ca. 0.9 h−1). More importantly, no tertiary amine was detected in the crude product. Thus, this method has the potential for the synthesis of multifunctional aniline derivatives from diversified fluoro-aromatic substrates (Scheme 5).
Zhang et al. [89] developed a method for the nickel-catalyzed coupling of fluoro-aromatics with oxazole. Studies showed that the heteroaryl fluoride is a class of substrate that can tolerate various functional groups. The mechanism of the C–F/C–H metathesis coupling reaction was thusly proposed: fluoro-aromatics first coordinate with Ni(0), I, and then C–F is oxidatively added to the center of Ni(0) to obtain Ni(II) intermediate II. In the presence of the base, the deprotonation reaction of benzoxazole generates a benzoxazole anion, which undergoes the metal transformation with intermediate II to form diaryl intermediate III. Finally, the required product is obtained through reductive elimination and regeneration of active Ni(0) species I (Scheme 6).

2.2. Pd-Catalyzed Activation of C–F Bonds in Fluoro-Aromatics

Other than nickel species, palladium complexes were also frequently utilized in the activation of C–F bonds.
Suzuki coupling is a heavily explored C–C formation reaction involving aryl groups [90,91,92]. Widdowson et al. carried out the first successful Suzuki coupling of fluoro-aromatics [93]; the ortho-substituted nitrofluorobenzene can proceed smoothly under Suzuki coupling conditions (TOF ca. 1.1 h−1). However, no coupling product is detected when the ortho-position is substituted by –CF3. Thus, the reaction clearly indicated that the ortho-nitro group is a necessary electron-withdrawing group for the catalytic activation of the C–F bond. Two possible reaction pathways may have operated in the products’ formation: cooperative insertion (path a) or addition-elimination sequence (path b) (Scheme 7).
Wang et al. reported a palladium-catalyzed cross-coupling between electron-deficient fluoro-aromatics and N-tosylhydrazones; the reaction can tolerate variety of functionalities on the substrates [94]. Using a substrate with a strong electron-withdrawing CF3 group, the coupling reaction proceeded smoothly with up to 70% yield (TOF ca. 1.8 h−1). Mechanistic study showed the reaction path includes activation of C–F bond and migration insertion of palladium carbene as two key steps, which are useful in preparing 1,1-diaryl ethylene derivatives (Scheme 8).
In Suzuki-type cross-coupling reactions, direct cross-coupling of organometallic reagents with fluoro-aromatics is relatively rare. In 2005, Dankwardt et al. carried out a microwave-assisted reaction between fluoro-aromatics and Grignard reagent [84] (Scheme 9).
Saeki et al. found that nickel and palladium complexes are complementary catalysts for the cross-coupling of Grignard reagent and C–F bond in fluoro-aromatics, but with different selectivity. The team used a DPPF-PdCl2 system in the cross-coupling of difluorobenzene with aryl Grignard reagents [83]. The chelating effect of the adjacent F-atom may promote the oxidative addition and facilitate the C–F bond activation. Thus, the Pd (1 mol%) catalyzed reaction of ortho-difluorobenzene with aryl Grignard reagent can obtain a single substituted product with up to 91% yield (TOF ca. 1.9 h−1), while the yield of cross-coupling products of meta-and para-difluorobenzene substrates is low. Subsequently, the study showed the reaction of 1,2,3-trifluorobenzene with Grignard reagent, using DPPP-NiCl2 (5 mol%) to obtain mainly the double coupling product with low conversion, while the DPPF-PdCl2 (1 mol%)-catalyzed reaction selectively received the single substitution product in high yields (Scheme 10).
Manabe et al. reported a process of ortho-selective cross-coupling of fluorobenzene with Grignard reagents, and PdCl2 (PCy3)2 was found to be an excellent catalyst for the reaction. Whereas, electron-donating groups such as hydroxyl, hydroxymethyl, and amino groups on the ortho-position of the fluoro-aromatics played a key role in accelerating the palladium-catalyzed cross-coupling [95] (TOF ca. 1.0–1.8 h−1, Scheme 11).
Cao et al. showed a palladium-catalyzed cross-coupling of fluoro-aromatics with B2pin2 in the presence of LiHMDS, and proposed an efficient method for the synthesis of aryl borate pinacol ester [96]. In general, fluoro-aromatics containing electron-neutral groups, weak electron-donating groups, or strong electron-donating groups have a higher yield than substrates containing electron-withdrawing groups. Under optimal conditions, the expected boronates are obtained in moderate to good yields (R = Ph, TOF ca. 3.9 h−1). In addition, the reaction tolerates wide range of functional groups, and no external ligand is needed. The plausible reaction pathway may be illustrated as such; at first, a boron atom of B2pin2 coordinates with a strong base LiN (SiMe3)2 to form a Lewis adduct of sp2-sp3 diborane species I. The B-B bond of the diborane species can be activated and then undergo heterolytic cleavage. The oxidative addition of the C-F bond in the fluoro-aromatics to the Pd(0) complex produces LArPd (II) F adduct . Subsequently, the trans-metallization of intermediates I and III formed LArPd(II)Bpin IV and intermediate II. Finally, the reductive elimination from IV provides the required aryl borate esters, accompanied by the regeneration of the active catalytic Pd(0) species (Scheme 12). The LiHMDS played the role of binding and breaking up the B2pin2 unit to facilitate the activation and transformation of the C–F bond of fluoro-aromatics.
Cao et al. created a catalytic system for the Sonogashira coupling of fluoro-aromatics with terminal alkynes, utilizing a Pd-catalyst. Thus, in the presence of LiHMDS, various electron-rich and/or electron-deficient fluoro-aromatics can be converted to aryl substituted alkynes (TOF ca. 2.2 h−1) [97]. The scope of this transformation can also be extended to aryl chlorides and bromides, and a plausible reaction mechanism may be illustrated as such. The first step involves the oxidative addition of C-F bonds of fluoro-aromatics to the Pd(0)L2 species to generate the intermediate L2ArPd(II)F complex I; subsequently, coordination of complex I with alkynes resulted in the formation of palladium alkynes transition state complex II; the coordinated alkynes are deprotonated by LiHMDS to produce complex ; and finally, the product and the active catalyst Pd(0)L2 species are generated by reductive elimination of intermediate III to complete the catalytic cycle (Scheme 13).
In general, methods for the selective activation of C–F bonds in multifluoro-aromatics to obtain partially fluoro-substituted aromatics are highly desirable. Zhang et al. developed a practical method for activating the C–F bonds of multifluoro-aromatics [98]. Thus, utilizing a palladium catalytic system, multifluorophenyl-pyridine substrates and triethylsilane undergo ortho-selective hydro-defluorination (HDF) with yields up to 98%. This method has the characteristics of low cost, wide substrate range, mild operation conditions, high efficiency, good regio-selectivity, etc. (Scheme 14).
Recently, Zhang et al. developed a method for the palladium-catalyzed arylation of multifluoro-aromatics [99]. The catalytic system, Pd(OAc)2/BrettPhos, is considered to be highly effective (TOF ca. 3.1 h−1), and has the advantages of high efficiency, wide substrate range, high regio-selectivity, and toleration of nitrogen-containing heterocyclic groups. In order to further improve the scope of the process, selective activation of multifluoro-aromatic substrates was carried out, and cross-coupling of trifluoro-aromatic with arylboronic acid in a gram scale reaction demonstrated fairly good results, which confirms the generality of this Pd-catalytic system. Preliminary mechanism studies have shown that the high regio-selectivity of the current reaction may be attributed to the electron-rich palladium complex Pd(0)BrettPhos, which may facilitate the oxidative addition of the C–F bonds. The catalytic system opens up a new scheme for the activation of C–F bonds in multifluoro-aromatics for the organic synthesis and related chemistry (Scheme 15).
Differing from the Zhang’s process [98], Grey and others studied Pd-catalyzed hydro-defluorination (HDF) of heterocyclic fluoro-aromatics (TOF ca. 1.3 h−1) [100] and extended the scope of this reaction to a range of hetero-aromatic scaffolds commonly encountered in pharmaceutical chemistry (Scheme 16).

2.3. Other Transition Metal Complexes Catalyzed C–F Activation of Fluoro-Aromatics

In comparison with nickel and palladium-catalyzed C–F bond activation, the same class of reaction, catalyzed with other transition metals, are less popular to date. In recent years, naturally abundant metals such as copper, cobalt, and iron have found increased application in the activation and functionalization of fluoro-aromatic C–F bonds.
Transition metal-mediated hydro-defluorination (HDF) of multifluoroaromatics is a fairly economic route for preparing partially fluorinated aromatics, and the methods’ development has received increased attention. The selective conversion of C–F bonds in multifluoro-aromatics is considered a valuable practice [101]. Zhang et al. created a Cu-catalyzed HDF of muitifluoro-aromatics (TOF ca. 2.7 h−1) [102]. Under optimal reaction conditions, pentafluoro-arenes with electron-withdrawing substituents (such as NO2, CN, and CF3) have fairly high reactivity, and the selective formation of tetrafluoroarene is also rapid (Scheme 17).
Crimmin et al. utilized a relatively rare example of Rh-mediated HDF of multifluoro-aromatics [103]; the catalytic system is highly selective, and the reaction directly activates C–F bonds adjacent to a C–H bond, with regio-selectivity up to 98.5–99% (Scheme 18).
Ding et al. also studied Rh-catalyzed HDF of heterocyclic multifluoro-aromatics via activation of C–F bonds in the ortho position of fluoro-aromatics [104]. Through the use of Rh-complexes and ethanol as hydride sources, the system can facilitate ortho-selective mono-HDF (TOF ca. 0.7 h−1) or double-HDF (TOF ca. 0.8 h−1) of multifluoro-aromatics. Mechanistic studies have shown that the phosphine ligands are crucial to catalytic performance, in which bidentate phosphine ligands are out-performing monodentate phosphine ligands in terms of product yields. In addition, phosphine ligands with higher steric hindrance are more favorable for producing mono-HDF products (Scheme 19).
Differing from the HDF reaction of multifluoro-aromatics, Iwasaki et al. showed the activation of aliphatic C–F bond catalyzed by cobalt, and realized the catalytic cross-coupling reaction of alkyl fluoride with Grignard reagent obtained the corresponding coupling products in fairly high yields [105]. Duan et al. also studied a co-catalyzed coupling reaction of fluoro-aromatics with Grignard reagents, but in the absence of phosphine or NHC ligands [65]. The co-mediated ‘easy-to-catalyze’ biaryl cross-coupling reaction through the cleavage of C–F bond is unexpected, in sharp contrast to the Ni or Pd catalyzed reactions, in which the existence of activated group(s) on fluoro-aromatics are necessary. However, the current C–F activation reaction is only catalyzed by CoCl2/DMPU; it is also noteworthy that in the coexisting of C–F, C–Cl and C–Br functionalities, highly selective C–F activation can be realized. Mechanistically, the assumption of the synergistic effect of Co–Ti bi-nuclear intermediate may play a key role in promoting the cleavage of the C–F bond. These findings will inspire further development of high-efficiency cobalt catalyst systems for C–F bond activation (Scheme 20).
Lee et al. carried out a co-catalyzed boration of C–F bonds in fluoro-aromatics [106]. This was the first time that co-catalyzed defluoro-boration of fluoro-aromatics was carried out in mild and practical conditions; it exhibited high selectivity for C–F bonds, even exceeding the boration on C–H bonds in the same substrates. This method makes it possible to direct the functionalization of a series of fluoro-aromatics. In addition, the catalytic system can tolerate unprotected functional groups (e.g., alcohols and amines), and can also activate C–F bonds under aerobic conditions, while somewhat sacrificing productivity (Scheme 21).
Based on the palladium-catalyzed activation of C–F bond of fluoro-aromatics promoted by LiHMDS, Cao et al. extended the catalytic system to iron [96] complexes. At present, though the activity (R = Ph, TOF ca. 0.3 h−1) of the catalyst is lower than that of Pd system, the Fe-catalyst has the advantages of low cost, low toxicity, environmentally friendliness and high availability (Scheme 22).
Recently, Nakamura et al. developed a new Fe-catalyzed ortho-C–F activation of diarylamine to synthesize DADHPs in one pot (TOF ca. 65–109 h−1) [107]. The Fe-catalytic system has good regio-selectivity, and can selectively synthesize DADHPs with different halogen substituents (fluorine, chlorine, and bromine). Increasing the structural diversity and availability of DADHPs, will help further development of functional molecules in the fields of materials science and synthetic chemistry (Scheme 23).

3. Activation of C–F Bond in Fluoro-Aromatics Promoted by Transition Metal-Free Processes

Most of the C–F bond activation in fluoro-aromatics is accomplished by transition metals [108,109,110,111]. However, the choice of different metal and/or ligands in the system will affect the outcoming products to somewhat unpredictable degrees. The activation of the C–F bond catalyzed by transition metal frequently requires harsh reaction conditions, and the transition metals may bring adverse effects to the working environment. Therefore, it is highly desirable to develop new, efficient and environmentally benign systems for chemical bond transformation reactions, and C–F bond activation/functionalization is one of the frontiers of inert chemical bonds activation [112]. In recent years, some efforts have also been made to explore the transition metal-free activation of C–F bonds in fluoro-aromatics.
Li et al. reported a simple method for activation of C–F bonds [113] without transition metal mediation. Thus, perfluoropyridine and Grignard reagent can undergo cross-coupling under ambient temperature, and results showed that perfluorinated aromatics can react with aryl Grignards in general; alkyl Grignards are also suitable for the cross coupling (Scheme 24).
Li et al. also showed the coupling reaction between multifluoro-aromatic imines and Grignard reagent without the participation of transition metal [114]. It was found that the electron-withdrawing effect of amino-functionality on the multifluoro-aromatic ring can weaken the bond energy of the ortho-C–F bond, which is beneficial to the activation of C–F bond in multifluoro-aromatic imines. This method is applicable to the coupling reaction between various fluoroarylimines and Grignard reagent, and can obtain ortho-substituted benzaldehyde derivatives with high yield (Scheme 25).
Cao et al. studied the metal-free cross-coupling of fluorinated phenyl pyridine with a variety of Grignard reagents [115]. The speculating mechanism indicates, at first, that the nitrogen atom on the pyridine ring coordinates with the magnesium ion of the Grignard reagent to form a magnesium complex. A; synergistic formation of a six-membered ring transition state, B; electrons’ rearrangement in the aromatic ring to form a complex C; and finally, C–F bond cleavage which forms the coupling products (Scheme 26).
In 2018, Cao et al. showed that the 3-arylation of indene can be directly and conveniently prepared via HMPA-promoted cleavage of the C–F bond of fluoro-arene in the presence of LDA [116], and that the reaction can be completed in 30 min at ambient temperature. A plausible mechanism is proposed. First, indene deprotonates in the presence of LDA to generate indenyl lithium I. In the presence of LDA, the cleavage of the C–F bond provides a key intermediate aryne II (Scheme 27); subsequently, I is added to intermediate II to obtain lithiated 1- or 3-arylalkenes (III or IV), intermediate III is isomerized under alkaline conditions to form more stable intermediate IV, and finally, the desired compound is obtained by neutralizing IV with water. In addition, the coordination of HMPA/THF with lithium ions has a significant effect on the reaction. It is expected that further expanding the scope of the reaction is highly possible.
In 2017, Ding et al. reported an economical and environmentally friendly cross-coupling reaction between fluoro-aromatics and amines, which involved N-heterocyclic-assisted selective C–F bond cleavage without the need for transition metal catalysts [112]. The reaction selectively cleaves the ortho-C–F bond of the difluorophenyl pyridine, while the para-C–F bond remains intact. The mechanistic studies showed that the cross-coupling reaction is promoted by intra-molecular Li/F interaction (Scheme 28).
Extending the scope of the heterocyclic-assisted C–F cleavage, Ding et al. created an efficient process for the activation of C–F bonds in polycyclic hetero-aromatics [117]. The reaction may proceed through the following route: lithium ions preferentially coordinate with a lone pair of polycyclic fluoro-aromatic nitrogen atoms, and through binding up the adjacent F atoms, the C–F bond energy is weakened. The incoming amine substrate nucleophilic attacks the carbon where the C–F is activated, and thereby forms the new C–N bond while simultaneously releasing LiF salt (Scheme 29).
It was found that multifluoro-benzenes can undergo a metathesis reaction with nucleophilic amines, and primary amines are easily reacted with multifluoro-aromatics to obtain aromatic amines [118] (Scheme 30).
For activation of C-F bond in multifluoro-aromatics, Tokárová et al. demonstrated consecutive nucleophilic substitutions of hexafluorobenzene and 1-pentafluorophenyl-1H-pyrrole (1a) with pyrrole/NaH and 2,5-dimethylpyrrole/NaH [119]. Results showed that the substitution mode of stepwise defluorinating specific fluorine atoms depends on the nature and quantity of nucleophiles used (pyrrole/NaH and 2,5-dimethylpyrrole/NaH). The reaction of hexafluorobenzene with pyrrolidinyl sodium salt (generated in situ from equimolar amounts of pyrrole and sodium hydride) proceeded smoothly to fully substituted compound 2 in 82% yield (Scheme 31).
Crimmin et al. reported a transition metal-free hydro-defluorination (HDF) process [120], and it was believed that the boron hydrides may represent a new tool for the activation of C-F bonds in transition metal-free HDF systems (Scheme 32).
Due mainly to its environmentally friendly and economical characteristics, as well as the possibility of utilization in some sophisticate synthetic practices, photo-catalysis also appears to be a useful tool in C–F bond activation/transformations [121]. Compared with the previous studies on the activation of C–F bond by cross-coupling of fluoro-aromatics with Grignard reagents, Zhang et al. demonstrated a selective mono-HDF of hexafluorobenzene, under transition metal-free photocatalysis conditions, to yield pentafluorobenzene in fairly high yields [122]. The photo-catalytic method can be applied to HDF of multifluorobenzenes. Research has shown that the steric hindrance of the photocatalyst and the multifluoro-aromatics largely determine the HDF rate, pointing to an internal sphere electron transfer pathway. The study emphasized the importance of the size and shape of photocatalyst and substrates in controlling the electron transfer mechanism and rate law. To further prove the potential of the transition metal-free photocatalysis, hexafluorobenzene was reacted with cyclohexene, where Py3 is used as a metal-free photocatalyst to generate hexafluorophenyl radicals, which are intercepted by 6.0 equivalents of cyclohexene, thus obtaining C–C coupling products with good yield. The study demonstrated the potential of obtaining partially fluorinated aromatics through photocatalytic HDF and the cross-coupling of multiflorophenyl with olefinic substrates (Scheme 33).
Wu et al. demonstrated a visible light-catalyzed defluoro-boration of multifluoro-aromatics with high selectivity [121]. Subsequently, Yang et al. also showed the photocatalytic boration of multifluoro-aromatics with NHC borane [123]. This transition metal-free photocatalytic process can directly generate B–H bonds in the aromatic products. With good functional group tolerance and high regioselectivity characteristics, the method provided nonparallel advantages for the preparation of a large number of valuable multifluoroaryl borane compounds, which further enriches the photocatalytic defluoroboration (DFB) of multifluoro-aromatics (Scheme 34).
Studer et al. showed that fluoro-aromatics undergo cross-metathesis defluoro-silylation, through synergistic nucleophilic aromatic substitution similar to the nickel-catalyzed cross-coupling of C–Si bond, and provided a synthetic method for obtaining aryl-substituted silanes through C–F bond activation [124]. Studies have shown that silicon-based lithium reagents such as PhMe2SiLi or Ph2t-BuSiLi, which are easily generated in situ from their hydrosilane analogs, react with various fluoro-aromatics to obtain corresponding highly substituted aromatic silanes. Compared with the classic nucleophilic aromatic substitutions, this transition metal-free, synergistic and nucleophilic aromatic substitution defluoro-silylation reaction ais lso suitable for substrates bearing relatively electron-rich substituents (Scheme 35).
Hua et al. demonstrated a nucleophilic substitution of fluoro-aromatics with various nucleophiles (such as alcohols, phenols, amines, amides, and N-heterocyclic compounds) [125]. The nucleophilic substitution uses KOH/DMSO as a medium, under mild, transition metal-free conditions, and provides an alternative alkali-promoted C–F bond activation process. Studies have shown that fluoro-aromatics with either electron-withdrawing or electron-donating groups can undergo smooth nucleophilic substitution, and have made the activation of the fluoro-aromatics’ C–F bond with electron donating functionalities, such as amide, bromine, cyano, aldehyde, acetyl, etc., possible. (Scheme 36).
In 2019, Deck et al. reported the nucleophilic activation of the C–F bond in ortho-fluoroaniline [126]. Fluorinated aniline reacts with stoichiometric Ti(NMe2)4 in 1,3,5-trimethylbenzene to obtain the ortho-defluoroamination products with good selectivity and yields (Scheme 37).
Shi et al. prepared a series of PNN-type tridentate coordination organic molecules through the metathesis of 2-diphenylphosphinyl arylamino-lithium salt with fluoro-aromatics [127]. This provided a transition metal-free protocol for the synthesis of PN-type tridentate ligands via activation of C–F bond in ortho-fluoroaniline. The method certainly enriched the PNN tridentate-ligated transition metal complexes toolbox, and the PNN tridentate-ligated transition metal catalytic chemistry (Scheme 38).

4. Conclusions and Outlook

Method development for the activation and conversion of inert C–F bonds in fluoro-aromatics is highly desirable for both routine organic synthesis and industrial and environmental applications. In recent years, effective and selective processes and systems developed for the activation and transformation of C–F bonds have seen remarkable signs of progress. However, in comparison with the activation of C–H and C–C bonds, the C–F bond activation fields still lack systematic concepts and theories.
As mentioned in this text, a good sign is that the activation and transformation of C–F bonds in fluoro-aromatics have seen increasing attention, and the advances made not only demonstrate new processes for the synthesis of novel fluorinated organic compounds, but also provide insight into theoretical perspectives, inspiring new concepts and methodologies in the search for more efficient C–F bond activation and transformation processes and systems.
The activation and functionalization of C–F bonds catalyzed by transition metal complexes are still the main focus of the subject. However, owing to the abnormal strength of the C-F bond energy and the shielding effects of the fluorine atom, currently, the transition metal-catalyzed sp2 C–F bonds activated in fluorinated aromatics are generally suffering from low catalytic efficiency, the high loading of catalyst (1–10 mol%) and the forced reaction conditions (e.g., the prolonged reaction time or high reaction temperature needed to ensure sufficient substrate conversion) as well the disappointing TOF numbers. Moreover, the precious metals (Pd, Pt, Rh, etc.)-catalyzed reactions showed no advantageous catalytic performances from an economic point of view, and the metal-catalyzed C–F activation may draw attention to the first-row transition metals for future developments toward practical applications.
On the other hand, economic and environmentally friendly chemical bond activation processes/methods are attracting increased attention. The development of low-cost, environmentally benign, highly efficient C–F activation and conversion, especially the transition metal-free processes, has recently opened up a window for the stochiometric reaction of fluorinated aromatics with a variety of reactive substrates; this should generate sustainable and sufficient interest to warrant further research works. Metal-free C–F bond activation processes have the advantage of easy scale-up and fewer environmental concerns; therefore, they will be one of the focal spots of C–F bond activation research.
In addition, ever since it was first demonstrated in 2006 that Frustrate Lewis Pairs (FLP) can split H–H bonds, FLP-catalyzed reactions are exhibiting outstanding performances in the activation of a wide variety of high-bonding-energy species [128,129,130], including organic molecules possessing sp3 and sp2 C–F bonds [68]. FLP-catalyzed C–F bond activation paves a new avenue for efficient and selective functionalization of multifluoro organic species to yield desirable new products. It is believed that FLP-catalyzed C–F bond activation will provide new chemical processes for practical applications.
It is expected that this paper will be of some help to researchers working in the C–F activation fields. The literature in this text is up to date at the end of 2022, and the fast development in the field will certainly bring some outcomes which you may find valuable. We will stay alert to the field to announce the most exciting research news periodically.
Together with the rapid development of organic chemistry, C–F bonds activation/functionalization will be one of the frontiers of research, and it is expected that the activation/functionalization of the C–F bond in fluoro-aromatics will keep its fast-developing pace. Thus, more fundamental and general processes/methods will be available for routine synthetic chemistry and application in industry.

Author Contributions

Conceptualization, writing-review and editing, supervision, funding acquisition, S.X.; original draft preparation, investigation R.M.; lab advisory and investigation, H.H. and X.L.; resources and project administration, G.M.; references validation, Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Nnational Science Foundation of China (NSFC) through project number 22172024.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Politanskaya, L.V.; Selivanova, G.A.; Panteleeva, E.V.; Tretyakov, E.V.; Platonov, V.E.; Nikul’shin, P.V.; Vinogradov, A.S.; Zonov, Y.V.; Karpov, V.M.; Mezhenkova, T.V. Organofluorine chemistry: Promising growth areas and challenges. Russ. Chem. Rev. 2019, 88, 425–569. [Google Scholar] [CrossRef]
  2. O’Hagan, D. Understanding organofluorine chemistry. An introduction to the C–F bond. Chem. Soc. Rev. 2008, 37, 308–319. [Google Scholar] [CrossRef] [PubMed]
  3. Tsui, G.C.; Hu, J. Organofluorine chemistry. Asian J. Org. Chem. 2019, 8, 566–567. [Google Scholar] [CrossRef]
  4. Uneyama, K. Organofluorine Chemistry; 9600 Garsington Road; Blackwell Publishing Ltd.: Oxford, UK, 2008; pp. 1–90. [Google Scholar]
  5. 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]
  6. Reichel, M.; Karaghiosoff, K. Reagents for selective fluoromethylation: A challenge in organofluorine chemistry. Angew. Chem. Int. Ed. 2020, 59, 12268–12281. [Google Scholar] [CrossRef]
  7. Meera, G.; Rohit, K.; Treesa, G.S.; Anilkumar, G. Advances and Prospects in Gold-Catalyzed C–H Activation. Asian. J. Org. Chem. 2020, 9, 144–161. [Google Scholar] [CrossRef]
  8. Bumberger, A.E.; Gordon, C.P.; Trummer, D.; Copéret, C. C–H Activation and Olefin Insertion in d8 and d0 Complexes: Same Elementary Steps, Different Electronics. Helv. Chim. Acta 2020, 103, e1900278. [Google Scholar] [CrossRef]
  9. Jin, L.; Yao, Q.-J.; Xie, P.-P.; Li, Y.; Zhan, B.-B.; Han, Y.-Q.; Hong, X.; Shi, B.-F. Atroposelective synthesis of axially chiral styrenes via an asymmetric C–H functionalization strategy. Chem 2020, 6, 497–511. [Google Scholar] [CrossRef]
  10. Liu, M.; Niu, J.-L.; Yang, D.; Song, M.-P. Development of a traceless directing group: Cp*-free cobalt-catalyzed C–H activation/annulations to access isoquinolinones. J. Org. Chem. 2020, 85, 4067–4078. [Google Scholar] [CrossRef]
  11. Palumbo, C.T.; Scopelliti, R.; Zivkovic, I.; Mazzanti, M. C–H bond activation by an isolated dinuclear U (III)/U (IV) nitride. J. Am. Chem. Soc. 2020, 142, 3149–3157. [Google Scholar] [CrossRef]
  12. Shi, Y.; Xing, H.; Huang, T.; Liu, X.; Chen, J.; Guo, X.; Li, G.-B.; Wu, Y. Divergent C–H activation synthesis of chalcones, quinolones and indoles. Chem. Commun. 2020, 56, 1585–1588. [Google Scholar] [CrossRef] [PubMed]
  13. Noor, A.; Qayyum, S.; Schwarz, S.; Dietel, T.; Kempe, R. Formation of a dimeric tungsten (i) complex via C–H activation. Dalton Trans. 2020, 49, 1992–1996. [Google Scholar] [CrossRef] [PubMed]
  14. Bhaskararao, B.; Singh, S.; Anand, M.; Verma, P.; Prakash, P.; Athira, C.; Malakar, S.; Schaefer, H.F.; Sunoj, R.B. Is silver a mere terminal oxidant in palladium catalyzed C–H bond activation reactions? Chem. Sci. 2020, 11, 208–216. [Google Scholar] [CrossRef] [Green Version]
  15. Hadlington, T.J.; Kostenko, A.; Driess, M. Cycloaddition Chemistry of a Silylene-Nickel Complex toward Organic π-Systems: From Reversibility to C–H Activation. Chem.-Eur. J. 2020, 26, 1958–1962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Fricke, C.; Dahiya, A.; Reid, W.B.; Schoenebeck, F. Gold-catalyzed C–H functionalization with aryl germanes. ACS Catal. 2019, 9, 9231–9236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Zhao, S.-N.; Wang, G.; Poelman, D.; Van Der Voort, P. Metal organic frameworks based materials for heterogeneous photocatalysis. Molecules 2018, 23, 2947. [Google Scholar] [CrossRef] [Green Version]
  18. Li, G.H.; Dong, D.Q.; Yang, Y.; Yu, X.Y.; Wang, Z.L. Direct Carbamoylation of Quinoline N-oxides with Hydrazinecarboxamides via C–H Bond Activation Catalyzed by Copper Catalyst. Adv. Synth. Catal. 2019, 361, 832–835. [Google Scholar]
  19. Ren, W.; Jin, M.; Zuo, Q.-M.; Yang, S.-D. Allylation of β-amino phosphonic acid precursor via palladium-NHC catalyzed allylic C–H activation. Org. Chem. Front. 2020, 7, 298–302. [Google Scholar] [CrossRef]
  20. Deng, L.; Dong, G. Carbon–carbon bond activation of ketones. Trends Chem. 2020, 2, 183–198. [Google Scholar] [CrossRef]
  21. Zhong, J.; Long, Y.; Yan, X.; He, S.; Ye, R.; Xiang, H.; Zhou, X. Rhodium-catalyzed pyridine N-oxide assisted Suzuki–Miyaura coupling reaction via C (O)–C bond activation. Org. Lett. 2019, 21, 9790–9794. [Google Scholar] [CrossRef]
  22. Nanda, T.; Ravikumar, P. A Palladium-Catalyzed Cascade C–C Activation of Cyclopropenone and Carbonylative Amination: Easy Access to Highly Functionalized Maleimide Derivatives. Org. Lett. 2020, 22, 1368–1374. [Google Scholar] [CrossRef] [PubMed]
  23. Ren, Y.; Lin, Z. Theoretical studies on Rh-catalyzed cycloisomerization of homopropargylallene-alkynes through C (sp3)–C (sp) bond activation. ACS Catal. 2020, 10, 1828–1837. [Google Scholar] [CrossRef]
  24. Long, Y.; Su, Z.; Zheng, Y.; He, S.; Zhong, J.; Xiang, H.; Zhou, X. Rhodium-catalyzed transarylation of benzamides: C–C bond vs C–N bond activation. ACS Catal. 2020, 10, 3398–3403. [Google Scholar] [CrossRef]
  25. Zhu, J.; Chen, P.-h.; Lu, G.; Liu, P.; Dong, G. Ruthenium-catalyzed reductive cleavage of unstrained aryl–aryl bonds: Reaction development and mechanistic study. J. Am. Chem. Soc. 2019, 141, 18630–18640. [Google Scholar] [CrossRef] [PubMed]
  26. Ambler, B.R.; Turnbull, B.W.; Suravarapu, S.R.; Uteuliyev, M.M.; Huynh, N.O.; Krische, M.J. Enantioselective Ruthenium-Catalyzed Benzocyclobutenone–Ketol Cycloaddition: Merging C–C Bond Activation and Transfer Hydrogenative Coupling for Type II Polyketide Construction. J. Am. Chem. Soc. 2018, 140, 9091–9094. [Google Scholar] [CrossRef] [PubMed]
  27. Avullala, T.; Asgari, P.; Hua, Y.; Bokka, A.; Ridlen, S.G.; Yum, K.; Dias, H.R.; Jeon, J. Umpolung α-Silylation of Cyclopropyl Acetates via Low-Temperature Catalytic C–C Activation. ACS Catal. 2018, 9, 402–408. [Google Scholar] [CrossRef] [PubMed]
  28. Lübbesmeyer, M.; Mackay, E.G.; Raycroft, M.A.; Elfert, J.; Pratt, D.A.; Studer, A. Base-Promoted C–C Bond Activation Enables Radical Allylation with Homoallylic Alcohols. J. Am. Chem. Soc. 2020, 142, 2609–2616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Meng, Q.; Wang, F.; Qian, P. Density functional calculations for Rh (I)-catalyzed C–C bond activation of siloxyvinylcyclopropanes and diazoesters. Appl. Organomet. Chem. 2019, 33, e4869. [Google Scholar] [CrossRef]
  30. Xia, Y.; Wang, J.; Dong, G. Suzuki–Miyaura coupling of simple ketones via activation of unstrained carbon–carbon bonds. J. Am. Chem. Soc. 2018, 140, 5347–5351. [Google Scholar] [CrossRef]
  31. Yu, C.-G.; Matsuo, Y. Nickel-catalyzed deaminative acylation of activated aliphatic amines with aromatic amides via C–N bond activation. Org. Lett. 2020, 22, 950–955. [Google Scholar] [CrossRef]
  32. Wang, Z.-X.; Yang, B. Chemical transformations of quaternary ammonium salts via C–N bond cleavage. Org. Biomol. Chem. 2020, 18, 1057–1072. [Google Scholar] [CrossRef] [PubMed]
  33. Kadam, A.A.; Metz, T.L.; Qian, Y.; Stanley, L.M. Ni-catalyzed three-component alkene carboacylation initiated by amide C–N bond activation. ACS Catal. 2019, 9, 5651–5656. [Google Scholar] [CrossRef]
  34. Li, C.-L.; Jiang, X.; Lu, L.-Q.; Xiao, W.-J.; Wu, X.-F. Cobalt (II)-catalyzed alkoxycarbonylation of aliphatic amines via C–N bond activation. Org. Lett. 2019, 21, 6919–6923. [Google Scholar] [CrossRef] [PubMed]
  35. Mishra, A.; Chauhan, S.; Verma, P.; Singh, S.; Srivastava, V. TBHP-Initiated Transamidation of Secondary Amides via C− N Bond Activation: A Metal-Free Approach. Asian. J. Org. Chem. 2019, 8, 853–857. [Google Scholar] [CrossRef]
  36. Wang, G.; Shi, Q.; Hu, W.; Chen, T.; Guo, Y.; Hu, Z.; Gong, M.; Guo, J.; Wei, D.; Fu, Z. Organocatalytic asymmetric N-sulfonyl amide CN bond activation to access axially chiral biaryl amino acids. Nat. Commun. 2020, 11, 946. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Jiang, X.; Zhang, M.M.; Xiong, W.; Lu, L.Q.; Xiao, W.J. Deaminative (Carbonylative) Alkyl-Heck-type Reactions Enabled by Photocatalytic C− N Bond Activation. Angew. Chem. Int. Ed. 2019, 58, 2402–2406. [Google Scholar] [CrossRef]
  38. Zhang, Y.; Yu, B.; Gao, B.; Zhang, T.; Huang, H. Triple-bond insertion triggers highly regioselective 1, 4-aminomethylamination of 1, 3-enynes with aminals enabled by Pd-catalyzed C–N bond activation. Org. Lett. 2019, 21, 535–539. [Google Scholar] [CrossRef]
  39. Zhou, T.; Ji, C.-L.; Hong, X.; Szostak, M. Palladium-catalyzed decarbonylative Suzuki–Miyaura cross-coupling of amides by carbon–nitrogen bond activation. Chem. Sci. 2019, 10, 9865–9871. [Google Scholar] [CrossRef] [Green Version]
  40. Wang, H.; Zhang, S.-Q.; Hong, X. Computational studies on Ni-catalyzed amide C–N bond activation. Chem. Commun. 2019, 55, 11330–11341. [Google Scholar] [CrossRef]
  41. Lang, S.M.; Bernhardt, T.M.; Bakker, J.M.; Yoon, B.; Landman, U. Methanol C–O Bond Activation by Free Gold Clusters Probed via Infrared Photodissociation Spectroscopy. Z. Phys. Chem. 2019, 233, 865–880. [Google Scholar] [CrossRef]
  42. Goulas, K.A.; Mironenko, A.V.; Jenness, G.R.; Mazal, T.; Vlachos, D.G. Fundamentals of C–O bond activation on metal oxide catalysts. Nat. Catal. 2019, 2, 269–276. [Google Scholar] [CrossRef]
  43. Khakyzadeh, V.; Rostami, A.; Veisi, H.; Shaghasemi, B.S.; Reimhult, E.; Luque, R.; Xia, Y.; Darvishi, S. Direct C–S bond formation via C–O bond activation of phenols in a crossover Pd/Cu dual-metal catalysis system. Org. Biomol. Chem. 2019, 17, 4491–4497. [Google Scholar] [CrossRef] [PubMed]
  44. Ma, H.; Bai, C.; Bao, Y.-S. Heterogeneous Suzuki–Miyaura coupling of heteroaryl ester via chemoselective C (acyl)–O bond activation. RSC Adv. 2019, 9, 17266–17272. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Tobisu, M.; Chatani, N. Nickel-catalyzed cross-coupling reactions of unreactive phenolic electrophiles via C–O bond activation. In Ni-and Fe-Based Cross-Coupling Reactions; Springer: Berlin/Heidelberg, Germany, 2017; pp. 129–156. [Google Scholar]
  46. Bisz, E.; Szostak, M. Iron-Catalyzed C− O Bond Activation: Opportunity for Sustainable Catalysis. ChemSusChem 2017, 10, 3964–3981. [Google Scholar] [CrossRef] [PubMed]
  47. Hoang, G.T.; Walsh, D.J.; McGarry, K.A.; Anderson, C.B.; Douglas, C.J. Development and mechanistic study of quinoline-directed acyl C–O bond activation and alkene oxyacylation reactions. J. Org. Chem. 2017, 82, 2972–2983. [Google Scholar] [CrossRef] [PubMed]
  48. Mondal, M.; Begum, T.; Bora, U. Chemoselective acyl C–O bond activation in esters for Suzuki–Miyaura coupling. Org. Chem. Front. 2017, 4, 1430–1434. [Google Scholar] [CrossRef]
  49. Malapit, C.A.; Caldwell, D.R.; Sassu, N.; Milbin, S.; Howell, A.R. Pd-catalyzed acyl C–O bond activation for selective ring-opening of α-methylene-β-lactones with amines. Org. Lett. 2017, 19, 1966–1969. [Google Scholar] [CrossRef] [Green Version]
  50. Purohit, P.; Seth, K.; Kumar, A.; Chakraborti, A.K. C–O Bond Activation by Nickel–Palladium Hetero-Bimetallic Nanoparticles for Suzuki–Miyaura Reaction of Bioactive Heterocycle-Tethered Sterically Hindered Aryl Carbonates. ACS Catal. 2017, 7, 2452–2457. [Google Scholar] [CrossRef]
  51. Wiensch, E.M.; Montgomery, J. Nickel-Catalyzed Amination of Silyloxyarenes through C–O Bond Activation. Angew. Chem. Int. Ed. 2018, 57, 11045–11049. [Google Scholar] [CrossRef] [PubMed]
  52. Lin, C.; Gao, F.; Shen, L. Advances in Transition Metal-Catalyzed Selective Functionalization of Inert C–O Bonds Assisted by Directing Groups. Adv. Synth. Catal. 2019, 361, 3915–3924. [Google Scholar] [CrossRef]
  53. Iyori, Y.; Takahashi, K.; Yamazaki, K.; Ano, Y.; Chatani, N. Nickel-catalyzed reductive defunctionalization of esters in the absence of an external reductant: Activation of C–O bonds. Chem. Commun. 2019, 55, 13610–13613. [Google Scholar] [CrossRef] [PubMed]
  54. Wilkinson, J.A. Recent advances in the selective formation of the carbon-fluorine bond. Chem. Rev. 1992, 92, 505–519. [Google Scholar] [CrossRef]
  55. Furuya, T.; Kuttruff, C.A.; Ritter, T. Carbon-fluorine bond formation. Curr. Opin. Drug Discov. Devel. 2008, 11, 803–819. [Google Scholar] [PubMed]
  56. Furuya, T.; Klein, J.E.; Ritter, T. Carbon-fluorine bond formation for the synthesis of aryl fluorides. Synthesis 2010, 2010, 1804–1821. [Google Scholar] [PubMed] [Green Version]
  57. Barthazy, P.; Hintermann, L.; Stoop, R.M.; Wörle, M.; Mezzetti, A.; Togni, A. Carbon–Fluorine Bond Formation via a Five-Coordinate Fluoro Complex of Ruthenium (II), Preliminary Communication. Helv. Chim. Acta 1999, 82, 2448–2453. [Google Scholar] [CrossRef]
  58. Campbell, M.G.; Ritter, T. Late-Stage Formation of Carbon–Fluorine Bonds. Chem. Rec. 2014, 14, 482–491. [Google Scholar] [CrossRef]
  59. Neumann, C.N.; Ritter, T. Facile C–F bond formation through a concerted nucleophilic aromatic substitution mediated by the phenofluor reagent. Acc. Chem. Res. 2017, 50, 2822–2833. [Google Scholar] [CrossRef]
  60. Pfeifer, L. New Methods and Reagents for Carbon-Fluorine Bond Formation; University of Oxford: Oxford, UK, 2016. [Google Scholar]
  61. Ye, Y.; Schimler, S.D.; Hanley, P.S.; Sanford, M.S. Cu (OTf) 2-mediated fluorination of aryltrifluoroborates with potassium fluoride. J. Am. Chem. Soc. 2013, 135, 16292–16295. [Google Scholar] [CrossRef]
  62. Park, N.H.; Senter, T.J.; Buchwald, S.L. Rapid synthesis of aryl fluorides in continuous flow through the Balz–Schiemann reaction. Angew. Chem. Int. Ed. 2016, 128, 12086–12090. [Google Scholar] [CrossRef]
  63. Thomas, H.P.; Wang, Y.-M.; Lorenzini, F.; Coles, S.J.; Horton, P.N.; Marr, A.C.; Saunders, G.C. Cyclometalation via carbon–fluorine bond activation induced by silver particles. Organometallics 2017, 36, 960–963. [Google Scholar] [CrossRef] [Green Version]
  64. Fujita, T.; Fuchibe, K.; Ichikawa, J. Transition-Metal-Mediated and-Catalyzed C–F Bond Activation by Fluorine Elimination. Angew. Chem. Int. Ed. 2019, 58, 390–402. [Google Scholar] [CrossRef]
  65. Wei, J.; Liu, K.-M.; Duan, X.-F. Cobalt-catalyzed biaryl couplings via C–F bond activation in the absence of phosphine or NHC ligands. J. Org. Chem. 2017, 82, 1291–1300. [Google Scholar] [CrossRef] [PubMed]
  66. Guo, Y.-Q.; Wang, R.; Song, H.; Liu, Y.; Wang, Q. Visible-light-induced deoxygenation/defluorination protocol for synthesis of γ, γ-difluoroallylic ketones. Org. Lett. 2020, 22, 709–713. [Google Scholar] [CrossRef] [PubMed]
  67. Yao, C.; Wang, S.; Norton, J.; Hammond, M. Catalyzing the hydrodefluorination of CF3-substituted alkenes by PhSiH3. H• transfer from a nickel hydride. J. Am. Chem. Soc. 2020, 142, 4793–4799. [Google Scholar] [CrossRef]
  68. Briceno-Strocchia, A.I.; Johnstone, T.C.; Stephan, D.W. Using frustrated Lewis pairs to explore C–F bond activation. Dalton Trans. 2020, 49, 1319–1324. [Google Scholar] [CrossRef] [PubMed]
  69. Ahmed, E.-A.M.; Suliman, A.M.; Gong, T.-J.; Fu, Y. Palladium-catalyzed stereoselective defluorination arylation/alkenylation/alkylation of gem-difluorinated cyclopropanes. Org. Lett. 2019, 21, 5645–5649. [Google Scholar] [CrossRef]
  70. Coates, G.; Tan, H.Y.; Kalff, C.; White, A.J.; Crimmin, M.R. Defluorosilylation of industrially relevant fluoroolefins using nucleophilic silicon reagents. Angew. Chem. Int. Ed. 2019, 58, 12514–12518. [Google Scholar] [CrossRef]
  71. Ding, D.; Lan, Y.; Lin, Z.; Wang, C. Synthesis of gem-difluoroalkenes by merging Ni-catalyzed C–F and C–C bond activation in cross-electrophile coupling. Org. Lett. 2019, 21, 2723–2730. [Google Scholar] [CrossRef]
  72. He, Y.; Anand, D.; Sun, Z.; Zhou, L. Visible-light-promoted redox neutral γ, γ-difluoroallylation of cycloketone oxime ethers with trifluoromethyl alkenes via C–C and C–F bond cleavage. Org. Lett. 2019, 21, 3769–3773. [Google Scholar] [CrossRef]
  73. Jiang, L.-F.; Ren, B.-T.; Li, B.; Zhang, G.-Y.; Peng, Y.; Guan, Z.-Y.; Deng, Q.-H. Nucleophilic substitution of gem-difluoroalkenes with TMSNu promoted by catalytic amounts of Cs2CO3. J. Org. Chem. 2019, 84, 6557–6564. [Google Scholar] [CrossRef]
  74. Kondoh, A.; Koda, K.; Terada, M. Organocatalytic Nucleophilic Substitution Reaction of gem-Difluoroalkenes with Ketene Silyl Acetals. Org. Lett. 2019, 21, 2277–2280. [Google Scholar] [CrossRef] [PubMed]
  75. Zeng, H.; Zhu, C.; Jiang, H. Single C (sp3)–F Bond Activation in a CF3 Group: Ipso-Defluorooxylation of (Trifluoromethyl) alkenes with Oximes. Org. Lett. 2019, 21, 1130–1133. [Google Scholar] [CrossRef] [PubMed]
  76. Yan, S.-S.; Wu, D.-S.; Ye, J.-H.; Gong, L.; Zeng, X.; Ran, C.-K.; Gui, Y.-Y.; Li, J.; Yu, D.-G. Copper-Catalyzed Carboxylation of C–F bonds with CO2. ACS Catal. 2019, 9, 6987–6992. [Google Scholar] [CrossRef]
  77. Ma, T.; Chen, Y.; Li, Y.; Ping, Y.; Kong, W. Nickel-catalyzed enantioselective reductive aryl fluoroalkenylation of alkenes. ACS Catal. 2019, 9, 9127–9133. [Google Scholar] [CrossRef]
  78. Kiplinger, J.L.; Richmond, T.G.; Osterberg, C.E. Activation of carbon-fluorine bonds by metal complexes. Chem. Rev. 1994, 94, 373–431. [Google Scholar] [CrossRef]
  79. Burdeniuc, J.; Jedicka, B.; Crabtree, R.H. Recent advances in C–F bond activation. Chem. Ber. 1997, 130, 145–154. [Google Scholar] [CrossRef]
  80. Torrens, H. Carbonfluorine bond activation by platinum group metal complexes. Coord. Chem. Rev. 2005, 249, 1957–1985. [Google Scholar] [CrossRef]
  81. Wilklow-Marnell, M.; Brennessel, W.W.; Jones, W.D. C (sp 2)–F Oxidative Addition of Fluorinated Aryl Ketones by iPrPCPIr. Organometallics 2017, 36, 3125–3134. [Google Scholar] [CrossRef]
  82. Mongin, F.; Mojovic, L.; Guillamet, B.; Trécourt, F.; Quéguiner, G. Cross-coupling reactions of phenylmagnesium halides with fluoroazines and fluorodiazines. J. Org. Chem. 2002, 67, 8991–8994. [Google Scholar] [CrossRef]
  83. Saeki, T.; Takashima, Y.; Tamao, K. Nickel-and palladium-catalyzed cross-coupling reaction of polyfluorinated arenes and alkenes with Grignard reagents. Synlett 2005, 2005, 1771–1774. [Google Scholar] [CrossRef]
  84. Dankwardt, J.W. Transition metal catalyzed cross-coupling of aryl Grignard reagents with aryl fluorides via Pd-or Ni-activation of the C–F bond: An efficient synthesis of unsymmetrical biaryls–application of microwave technology in ligand and catalyst screening. J. Organomet. Chem. 2005, 690, 932–938. [Google Scholar] [CrossRef]
  85. Xiao, S.-H.; Xiong, Y.; Zhang, X.-X.; Cao, S. Nickel-catalyzed N-heterocycle-directed cross-coupling of fluorinated arenes with organozinc reagents. Tetrahedron 2014, 70, 4405–4411. [Google Scholar] [CrossRef]
  86. He, Y.; Chen, Z.; He, C.Y.; Zhang, X. Nickel-Catalyzed ortho-Selective Hydrodefluorination of N-Containing Heterocycle-Polyfluoroarenes. Chin. J. Chem. 2013, 31, 873–877. [Google Scholar] [CrossRef]
  87. Cui, B.; Jia, S.; Tokunaga, E.; Shibata, N. Defluorosilylation of fluoroarenes and fluoroalkanes. Nat. Commun. 2018, 9, 4393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Harada, T.; Ueda, Y.; Iwai, T.; Sawamura, M. Nickel-catalyzed amination of aryl fluorides with primary amines. Chem. Commun. 2018, 54, 1718–1721. [Google Scholar] [CrossRef]
  89. Yin, Y.; Yue, X.; Zhong, Q.; Jiang, H.; Bai, R.; Lan, Y.; Zhang, H. Ni-Catalyzed C–F Bond Functionalization of Unactivated Aryl Fluorides and Corresponding Coupling with Oxazoles. Adv. Synth. Catal. 2018, 360, 1639–1643. [Google Scholar] [CrossRef]
  90. Kim, Y.M.; Yu, S. Palladium (0)-catalyzed amination, Stille coupling, and Suzuki coupling of electron-deficient aryl fluorides. J. Am. Chem. Soc. 2003, 125, 1696–1697. [Google Scholar] [CrossRef]
  91. Albéniz, A.C.; Casares, J.A. Palladium-mediated organofluorine chemistry. Adv. Organomet. Chem. 2014, 62, 1–110. [Google Scholar]
  92. Dehury, N.; Maity, N.; Tripathy, S.K.; Basset, J.-M.; Patra, S. Dinuclear tetrapyrazolyl palladium complexes exhibiting facile tandem transfer hydrogenation/Suzuki coupling reaction of fluoroarylketone. ACS Catal. 2016, 6, 5535–5540. [Google Scholar] [CrossRef] [Green Version]
  93. Widdowson, D.A.; Wilhelm, R. Palladium catalysed Suzuki reactions of fluoroarenes. Chem. Commun. 2003, 578–579. [Google Scholar] [CrossRef]
  94. Luo, H.; Wu, G.; Xu, S.; Wang, K.; Wu, C.; Zhang, Y.; Wang, J. Palladium-catalyzed cross-coupling of aryl fluorides with N-tosylhydrazones via C–F bond activation. Chem. Commun. 2015, 51, 13321–13323. [Google Scholar] [CrossRef] [PubMed]
  95. Manabe, K.; Ishikawa, S. Ortho-selective cross-coupling of fluorobenzenes with Grignard reagents: Acceleration by electron-donating ortho-directing groups. Synthesis 2008, 2008, 2645–2649. [Google Scholar] [CrossRef]
  96. Zhao, X.; Wu, M.; Liu, Y.; Cao, S. LiHMDS-promoted palladium or iron-catalyzed ipso-defluoroborylation of aryl fluorides. Org. Lett. 2018, 20, 5564–5568. [Google Scholar] [CrossRef] [PubMed]
  97. He, J.; Yang, K.; Zhao, J.; Cao, S. LiHMDS-promoted palladium-catalyzed Sonogashira cross-coupling of aryl fluorides with terminal alkynes. Org. Lett. 2019, 21, 9714–9718. [Google Scholar] [CrossRef]
  98. Chen, Z.; He, C.Y.; Yin, Z.; Chen, L.; He, Y.; Zhang, X. Palladium-Catalyzed Ortho-Selective C-F Activation of Polyfluoroarenes with Triethylsilane: A Facile Access to Partially Fluorinated Aromatics. Angew. Chem. Int. Ed. 2013, 52, 5813–5817. [Google Scholar] [CrossRef]
  99. Luo, Z.-J.; Zhao, H.-Y.; Zhang, X. Highly selective Pd-catalyzed direct C–F bond arylation of polyfluoroarenes. Org. Lett. 2018, 20, 2543–2546. [Google Scholar] [CrossRef]
  100. Gair, J.J.; Grey, R.L.; Giroux, S.; Brodney, M.A. Palladium catalyzed hydrodefluorination of fluoro-(hetero) arenes. Org. Lett. 2019, 21, 2482–2487. [Google Scholar] [CrossRef]
  101. Matsunami, A.; Kayaki, Y.; Kuwata, S.; Ikariya, T. Nucleophilic Aromatic Substitution in Hydrodefluorination Exemplified by Hydridoiridium (III) Complexes with Fluorinated Phenylsulfonyl-1, 2-diphenylethylenediamine Ligands. Organometallics 2018, 37, 1958–1969. [Google Scholar] [CrossRef]
  102. Lv, H.; Cai, Y.B.; Zhang, J.L. Copper-Catalyzed Hydrodefluorination of Fluoroarenes by Copper Hydride Intermediates. Angew. Chem. Int. Ed. 2013, 52, 3203–3207. [Google Scholar] [CrossRef]
  103. Ekkert, O.; Strudley, S.D.; Rozenfeld, A.; White, A.J.; Crimmin, M.R. Rhodium Catalyzed, Carbon–Hydrogen Bond Directed Hydrodefluorination of Fluoroarenes. Organometallics 2014, 33, 7027–7030. [Google Scholar] [CrossRef]
  104. Chen, J.; Huang, D.; Ding, Y. Rhodium-Catalyzed ortho-Selective C-F Activation and Hydrodefluorination of Heterocycle-Substituted Polyfluoroarenes: Dominated by Phosphine Ligands. Chem. Sel. 2017, 2, 1219–1224. [Google Scholar]
  105. Iwasaki, T.; Yamashita, K.; Kuniyasu, H.; Kambe, N. Co-catalyzed cross-coupling reaction of alkyl fluorides with alkyl Grignard reagents. Org. Lett. 2017, 19, 3691–3694. [Google Scholar] [CrossRef]
  106. Lim, S.; Song, D.; Jeon, S.; Kim, Y.; Kim, H.; Lee, S.; Cho, H.; Lee, B.C.; Kim, S.E.; Kim, K. Cobalt-Catalyzed C–F Bond Borylation of Aryl Fluorides. Org. Lett. 2018, 20, 7249–7252. [Google Scholar] [CrossRef] [PubMed]
  107. Aoki, Y.; O’Brien, H.M.; Kawasaki, H.; Takaya, H.; Nakamura, M. Ligand-Free Iron-Catalyzed C–F Amination of Diarylamines: A One-Pot Regioselective Synthesis of Diaryl Dihydrophenazines. Org. Lett. 2019, 21, 461–464. [Google Scholar] [CrossRef]
  108. Kuehnel, M.F.; Lentz, D.; Braun, T. Synthesis of fluorinated building blocks by transition-metal-mediated hydrodefluorination reactions. Angew. Chem. Int. Ed. 2013, 52, 3328–3348. [Google Scholar] [CrossRef] [PubMed]
  109. Fan, F.; Zhao, L.; Luo, M.; Zeng, X. Chromium-Catalyzed Selective Cross-Electrophile Coupling between Unactivated C (aryl)–F and C (aryl)–O Bonds. Organometallics 2022, 41, 561–568. [Google Scholar] [CrossRef]
  110. Matsunami, A.; Kuwata, S.; Kayaki, Y. Regioselective Transfer Hydrogenative Defluorination of Polyfluoroarenes Catalyzed by Bifunctional Azairidacycle. Organics 2022, 3, 150–160. [Google Scholar] [CrossRef]
  111. Ahrens, T.; Kohlmann, J.; Ahrens, M.; Braun, T. Functionalization of fluorinated molecules by transition-metal-mediated C–F bond activation to access fluorinated building blocks. Chem. Rev. 2015, 115, 931–972. [Google Scholar] [CrossRef]
  112. Chen, J.; Huang, D.; Ding, Y. C–N Bond Coupling of Fluorobenzenes: N-Heterocycle-Assisted Selective C–F Bond Cleavage through an Li/F Interaction. Eur. J. Org. Chem. 2017, 2017, 4300–4304. [Google Scholar] [CrossRef]
  113. Sun, Y.; Sun, H.; Jia, J.; Du, A.; Li, X. Transition-Metal-Free Synthesis of Fluorinated Arenes from Perfluorinated Arenes Coupled with Grignard Reagents. Organometallics 2014, 33, 1079–1081. [Google Scholar] [CrossRef]
  114. Lu, F.; Sun, H.; Du, A.; Feng, L.; Li, X. Selective alkylation and arylation of C–F bond with Grignard reagents. Org. Lett. 2014, 16, 772–775. [Google Scholar] [CrossRef]
  115. Xiong, Y.; Wu, J.; Xiao, S.; Xiao, J.; Cao, S. Noncatalytic pyridyl-directed alkylation and arylation carbon–fluorine bond of polyfluoroarenes with Grignard reagents. J. Org. Chem. 2013, 78, 4599–4603. [Google Scholar] [CrossRef]
  116. Ji, X.; Li, J.; Wu, M.; Cao, S. Facile Synthesis of 3-Arylindenes by HMPA-Promoted Direct Arylation of Indenes with Aryl Fluorides. ACS Omega 2018, 3, 10099–10106. [Google Scholar] [CrossRef] [PubMed]
  117. Chen, J.; Huang, D.; Ding, Y. Transition-metal-free site-selective C–F bond activation for synthesis of 8-aminoquinolines. Tetrahedron Lett. 2017, 58, 4240–4242. [Google Scholar] [CrossRef]
  118. Platonov, V.E.; Haas, A.; Schelvis, M.; Lieb, M.; Dvornikova, K.V.; Osina, O.I.; Gatilov, Y.V. Polyfluorinated arylnitramines. J. Fluor. Chem. 2001, 109, 131–139. [Google Scholar] [CrossRef]
  119. Tokárová, Z.; Balogh, R.; Tisovský, P.; Hrnčariková, K.; Vegh, D. Direct nucleophilic substitution of polyfluorobenzenes with pyrrole and 2,5-dimethylpyrrole. J. Fluor. Chem. 2017, 204, 59–64. [Google Scholar] [CrossRef]
  120. Phillips, N.A.; O’Hanlon, J.; Hooper, T.N.; White, A.J.; Crimmin, M.R. Dihydridoboranes: Selective Reagents for Hydroboration and Hydrodefluorination. Org. Lett. 2019, 21, 7289–7293. [Google Scholar] [CrossRef] [PubMed]
  121. Xu, W.; Jiang, H.; Leng, J.; Ong, H.-W.; Wu, J. Visible-Light-Induced Selective Defluoroborylation of Polyfluoroarenes, gem-Difluoroalkenes, and Trifluoromethylalkenes. Angew. Chem. Int. Ed. 2020, 132, 4038–4045. [Google Scholar] [CrossRef]
  122. Lu, J.; Khetrapal, N.S.; Johnson, J.A.; Zeng, X.C.; Zhang, J. “π-Hole– π” Interaction Promoted Photocatalytic Hydrodefluorination via Inner-Sphere Electron Transfer. J. Am. Chem. Soc. 2016, 138, 15805–15808. [Google Scholar] [CrossRef] [PubMed]
  123. Xia, P.-J.; Ye, Z.-P.; Hu, Y.-Z.; Xiao, J.-A.; Chen, K.; Xiang, H.-Y.; Chen, X.-Q.; Yang, H. Photocatalytic C–F Bond Borylation of Polyfluoroarenes with NHC-boranes. Org. Lett. 2020, 22, 1742–1747. [Google Scholar] [CrossRef]
  124. Mallick, S.; Xu, P.; Würthwein, E.U.; Studer, A. Silyldefluorination of Fluoroarenes by Concerted Nucleophilic Aromatic Substitution. Angew. Chem. Int. Ed. 2019, 131, 289–293. [Google Scholar] [CrossRef]
  125. Su, J.; Chen, Q.; Lu, L.; Ma, Y.; Auyoung, G.H.L.; Hua, R. Base-promoted nucleophilic fluoroarenes substitution of C-F bonds. Tetrahedron 2018, 74, 303–307. [Google Scholar] [CrossRef]
  126. Hough, S.E.; Hargrove, W.R., Jr.; Deck, P.A. Regioselective, nucleophilic activation of C-F bonds in o-fluoroanilines. J. Fluorine Chem. 2019, 224, 121–126. [Google Scholar] [CrossRef]
  127. Liu, H.; Yuan, H.; Shi, X. Synthesis of nickel and palladium complexes with diarylamido-based unsymmetrical pincer ligands and application for norbornene polymerization. Dalton Trans. 2019, 48, 609–617. [Google Scholar] [CrossRef] [PubMed]
  128. Slootweg, J.C.; Jupp, A.R. (Eds.) Frustrated Lewis Pairs. In Molecular Catalysis; Springer: Cham, Switzerland, 2021; Volume 2. [Google Scholar]
  129. Stephan, D.W. Frustrated Lewis Pairs: From Concept to Catalysis. Acc. Chem. Res. 2015, 48, 306–316. [Google Scholar] [CrossRef] [PubMed]
  130. Stephan, D.W.; Erker, G. Frustrated Lewis Pairs: Metal-free Hydrogen Activation and More. Angew. Chem. Int. Ed. 2010, 49, 46–76. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Activation of C–F Bond catalyzed by nickel complexes with bidentate phosphine ligand.
Scheme 1. Activation of C–F Bond catalyzed by nickel complexes with bidentate phosphine ligand.
Catalysts 12 01665 sch001
Scheme 2. NiCl2 (dppp) catalyzed cross–coupling of fluoro–aromatics with organozinc reagents.
Scheme 2. NiCl2 (dppp) catalyzed cross–coupling of fluoro–aromatics with organozinc reagents.
Catalysts 12 01665 sch002
Scheme 3. NiCl2 Catalyzed hydro–defluorination of fluoro–aromatics.
Scheme 3. NiCl2 Catalyzed hydro–defluorination of fluoro–aromatics.
Catalysts 12 01665 sch003
Scheme 4. Nickel(0)–catalyzed defluoro–silylation of fluoro–aromatics.
Scheme 4. Nickel(0)–catalyzed defluoro–silylation of fluoro–aromatics.
Catalysts 12 01665 sch004
Scheme 5. Nickel–catalyzed amination of fluoro–aromatics with primary amines.
Scheme 5. Nickel–catalyzed amination of fluoro–aromatics with primary amines.
Catalysts 12 01665 sch005
Scheme 6. Nickel–catalyzed cross–coupling of fluoro–aromatics with oxazole and the mechanistic pathway.
Scheme 6. Nickel–catalyzed cross–coupling of fluoro–aromatics with oxazole and the mechanistic pathway.
Catalysts 12 01665 sch006
Scheme 7. Palladium–catalyzed C–F bond activation of nitrofluoro–aromatics and the reaction mechanism.
Scheme 7. Palladium–catalyzed C–F bond activation of nitrofluoro–aromatics and the reaction mechanism.
Catalysts 12 01665 sch007
Scheme 8. Pd–catalyzed cross–coupling of electron–deficient fluoro–aromatics with N–tosylhydrazones and the reaction pathway.
Scheme 8. Pd–catalyzed cross–coupling of electron–deficient fluoro–aromatics with N–tosylhydrazones and the reaction pathway.
Catalysts 12 01665 sch008
Scheme 9. Palladium–catalyzed cross–coupling of fluoro-aromatics with PhMgCl under microwave assistance.
Scheme 9. Palladium–catalyzed cross–coupling of fluoro-aromatics with PhMgCl under microwave assistance.
Catalysts 12 01665 sch009
Scheme 10. Cross–coupling of fluoro–aromatic with Grignard reagents.
Scheme 10. Cross–coupling of fluoro–aromatic with Grignard reagents.
Catalysts 12 01665 sch010
Scheme 11. Cross–coupling of fluoro–aromatics with Grignard reagent catalyzed by PdCl2(PCy3)2.
Scheme 11. Cross–coupling of fluoro–aromatics with Grignard reagent catalyzed by PdCl2(PCy3)2.
Catalysts 12 01665 sch011
Scheme 12. LiHMDS–promoted palladium catalyzed defluoroborylation of fluoro–aromatics and reaction pathways.
Scheme 12. LiHMDS–promoted palladium catalyzed defluoroborylation of fluoro–aromatics and reaction pathways.
Catalysts 12 01665 sch012
Scheme 13. Plausible mechanism of LiHMDS–promoted Sonogashira cross–coupling of fluoro–ar–omatics with terminal alkynes catalyzed with palladium.
Scheme 13. Plausible mechanism of LiHMDS–promoted Sonogashira cross–coupling of fluoro–ar–omatics with terminal alkynes catalyzed with palladium.
Catalysts 12 01665 sch013
Scheme 14. Pd–catalyzed HDF of fluorophenyl–pyridine substrates with triethylsilane.
Scheme 14. Pd–catalyzed HDF of fluorophenyl–pyridine substrates with triethylsilane.
Catalysts 12 01665 sch014
Scheme 15. Pd(OAc)2/BrettPhos–catalyzed C–F bond activation of multifluoro–aromatics and plau–sible reaction pathway.
Scheme 15. Pd(OAc)2/BrettPhos–catalyzed C–F bond activation of multifluoro–aromatics and plau–sible reaction pathway.
Catalysts 12 01665 sch015
Scheme 16. Palladium–catalyzed HDF of heterocyclic fluoro–aromatics.
Scheme 16. Palladium–catalyzed HDF of heterocyclic fluoro–aromatics.
Catalysts 12 01665 sch016
Scheme 17. Cu–catalyzed HDF of pentafluoroaromatics.
Scheme 17. Cu–catalyzed HDF of pentafluoroaromatics.
Catalysts 12 01665 sch017
Scheme 18. Rh–catalyzed HDF of multifluoro–aromatics (BDIAlH2, hydrocarbon-soluble aluminum dihydride).
Scheme 18. Rh–catalyzed HDF of multifluoro–aromatics (BDIAlH2, hydrocarbon-soluble aluminum dihydride).
Catalysts 12 01665 sch018
Scheme 19. Rh–catalyzed HDF of heterocyclic substituted multifluoro–aromatics and selectivity pathways for phosphine ligand control.
Scheme 19. Rh–catalyzed HDF of heterocyclic substituted multifluoro–aromatics and selectivity pathways for phosphine ligand control.
Catalysts 12 01665 sch019
Scheme 20. Cobalt–catalyzed C–F bond activation and Co–Ti co–oxidative addition reaction transition state.
Scheme 20. Cobalt–catalyzed C–F bond activation and Co–Ti co–oxidative addition reaction transition state.
Catalysts 12 01665 sch020
Scheme 21. Co–catalyzed boration of fluoro–aromatics.
Scheme 21. Co–catalyzed boration of fluoro–aromatics.
Catalysts 12 01665 sch021
Scheme 22. LiHMDS–promoted Fe–catalyzed defluoroboration of fluoro–aromatics.
Scheme 22. LiHMDS–promoted Fe–catalyzed defluoroboration of fluoro–aromatics.
Catalysts 12 01665 sch022
Scheme 23. Ligand–free Fe–catalyzed C–F amination of diarylamine.
Scheme 23. Ligand–free Fe–catalyzed C–F amination of diarylamine.
Catalysts 12 01665 sch023
Scheme 24. Cross–coupling of perfluoropyridine and Grignard reagents.
Scheme 24. Cross–coupling of perfluoropyridine and Grignard reagents.
Catalysts 12 01665 sch024
Scheme 25. Cross–coupling between multifluoroarylimine and Grignard reagent while excluding transition metal.
Scheme 25. Cross–coupling between multifluoroarylimine and Grignard reagent while excluding transition metal.
Catalysts 12 01665 sch025
Scheme 26. The cross–coupling of multifluoro–arenes with Grignard reagents via pyridine–directed cleavage of the C–F bond in the absence of transition metal.
Scheme 26. The cross–coupling of multifluoro–arenes with Grignard reagents via pyridine–directed cleavage of the C–F bond in the absence of transition metal.
Catalysts 12 01665 sch026
Scheme 27. HMPA–promoted direct arylation of indenes with aryl fluorides and plausible reaction pathway.
Scheme 27. HMPA–promoted direct arylation of indenes with aryl fluorides and plausible reaction pathway.
Catalysts 12 01665 sch027
Scheme 28. Cross–coupling of fluoro–aromatics with amines under transition metal–free conditions.
Scheme 28. Cross–coupling of fluoro–aromatics with amines under transition metal–free conditions.
Catalysts 12 01665 sch028
Scheme 29. Cross–coupling of heterocyclic fluoro–aromatics with amines under transition metal–free conditions.
Scheme 29. Cross–coupling of heterocyclic fluoro–aromatics with amines under transition metal–free conditions.
Catalysts 12 01665 sch029
Scheme 30. Reaction of perfluorobenzene C–F bond with primary amines.
Scheme 30. Reaction of perfluorobenzene C–F bond with primary amines.
Catalysts 12 01665 sch030
Scheme 31. Nucleophilic substitution of multifluoro–aromatics with pyrrole/NaH and 2,5–Dime–thylpyrrole/NaH.
Scheme 31. Nucleophilic substitution of multifluoro–aromatics with pyrrole/NaH and 2,5–Dime–thylpyrrole/NaH.
Catalysts 12 01665 sch031
Scheme 32. Regio–selective HDF of fluoro–aromatics with dihydroboranes.
Scheme 32. Regio–selective HDF of fluoro–aromatics with dihydroboranes.
Catalysts 12 01665 sch032
Scheme 33. Photocatalystic HDF of perfluorobenzene and cross–coupling of multifluorobenzene with olefinc substrate.
Scheme 33. Photocatalystic HDF of perfluorobenzene and cross–coupling of multifluorobenzene with olefinc substrate.
Catalysts 12 01665 sch033
Scheme 34. Photocatalytic defluoro–boration of multifluoro–aromatics with NHC borane.
Scheme 34. Photocatalytic defluoro–boration of multifluoro–aromatics with NHC borane.
Catalysts 12 01665 sch034
Scheme 35. Transition metal–free defluoro–silylation of fluoro–aromatics.
Scheme 35. Transition metal–free defluoro–silylation of fluoro–aromatics.
Catalysts 12 01665 sch035
Scheme 36. Transition metal–free nucleophilic substitution of fluoro–aromatics with various nucle–ophiles.
Scheme 36. Transition metal–free nucleophilic substitution of fluoro–aromatics with various nucle–ophiles.
Catalysts 12 01665 sch036
Scheme 37. Activation of C–F bond in ortho–fluoroaniline.
Scheme 37. Activation of C–F bond in ortho–fluoroaniline.
Catalysts 12 01665 sch037
Scheme 38. Synthesis of PNN type tridentate ligands.
Scheme 38. Synthesis of PNN type tridentate ligands.
Catalysts 12 01665 sch038
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ma, R.; Hu, H.; Li, X.; Mao, G.; Song, Y.; Xin, S. Advances in Catalytic C–F Bond Activation and Transformation of Aromatic Fluorides. Catalysts 2022, 12, 1665. https://doi.org/10.3390/catal12121665

AMA Style

Ma R, Hu H, Li X, Mao G, Song Y, Xin S. Advances in Catalytic C–F Bond Activation and Transformation of Aromatic Fluorides. Catalysts. 2022; 12(12):1665. https://doi.org/10.3390/catal12121665

Chicago/Turabian Style

Ma, Rongqing, Hongfan Hu, Xinle Li, Guoliang Mao, Yuming Song, and Shixuan Xin. 2022. "Advances in Catalytic C–F Bond Activation and Transformation of Aromatic Fluorides" Catalysts 12, no. 12: 1665. https://doi.org/10.3390/catal12121665

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