Bo-Jie
Li†
a,
Yu-Long
Ruan†
b,
Lei
Zhu
*a,
Jian
Zhou
*b and
Jin-Sheng
Yu
*bc
aHubei Engineering University, Xiaogan, China. E-mail: Lei.zhu@hbeu.edu.cn
bState Key Laboratory of Petroleum Molecular & Process Engineering, Shanghai Engineering Research Center of Molecular Therapeutics and New Drug Development; School of Chemistry and Molecular Engineering, East China Normal University, Shanghai, 200062, P. R. China. E-mail: jzhou@chem.ecnu.edu.cn; jsyu@chem.ecnu.edu.cn
cKey Laboratory of Tropical Medicinal Resource Chemistry of Ministry of Education, Hainan Normal University, Haikou 571158, P. R. China
First published on 29th August 2024
Chiral organofluorine compounds featuring a monofluoromethyl (CH2F)-substituted stereocenter are often encountered in a number of drugs and bioactive molecules. Consequently, the development of catalytic asymmetric methods for the enantioselective construction of CH2F-substituted stereocenters has made great progress over the past two decades, and a variety of enantioselective transformations have been accordingly established. According to the types of fluorinated reagents or substrates employed, these protocols can be divided into the following major categories: (i) enantioselective ring opening of epoxides or azetidinium salts by fluoride anions; (ii) asymmetric monofluoromethylation with 1-fluorobis(phenylsulfonyl)methane; (iii) asymmetric fluorocyclization of functionalized alkenes with Selectfluor; and (iv) asymmetric transformations involving α-CH2F ketones, α-CH2F alkenes, or other CH2F-containing substrates. This feature article aims to summarize these recent advances and discusses the possible reaction mechanisms, advantages and limitations of each protocol and their applications. Synthetic opportunities still open for further development are illustrated as well. This review article will be an inspiration for researchers engaged in asymmetric catalysis, organofluorine chemistry, and medicinal chemistry.
Among them, monofluoromethyl-containing compounds have recently found increasing applications in drug and agrochemical discovery,2 because the monofluoromethyl (CH2F) group proves to be an ideal bioisoster of methyl, hydroxymethyl, and other functionalities,7 which often significantly improve the biological properties of parent molecules once introduced.5 In particular, CH2F-substituted stereocenters are widely present in drugs and bioactive molecules,8–12 as illustrated in Fig. 1. For example, carmegliptin (1), bearing a CH2F-substituted stereocenter, is an investigational oral anti-hyperglycemic drug for the treatment of type 2 diabetes;8 CH2F-containing nucleoside SFDC (3) has high antiherpes and antitumor activities;9 florfenicol (4) is a broad-spectrum bacteriostatic antibiotic used exclusively in veterinary medicine to treat the pathology of farm and aquatic animals;10 and (S)-α-CH2F-dopa (5) is an inhibitor of dopa decarboxylase.11 In addition, the use of monofluoromethyl to replace the methyl group of (S)-ibuprofen significantly improves its inhibitory activity toward COX-1 with almost an equal pharmacokinetic profile, thus showing enhanced analgesic activity and reduced gastric damage.12
In this context, the development of catalytic asymmetric methods for the diverse construction of CH2F-substituted stereocenters has gained much attention and has become a hot research topic. Accordingly, plenty of elegant, highly enantioselective catalytic transformations have been established to access various molecules featuring CH2F-substituted stereocenters in the past two decades. As demonstrated in Fig. 2, these protocols can be classified into the following major categories, according to the types of fluorinated reagents or substrates used in the reaction: (1) enantioselective ring opening of epoxides or azetidinium salts by fluoride anions; (2) asymmetric monofluoromethylation with 1-fluorobis(phenylsulfonyl)methane (FBSM); (3) asymmetric fluorocyclization of functionalized alkenes with Selectfluor; (4) asymmetric reduction or nucleophilic addition of α-CH2F ketones; (5) asymmetric transformations involving α-CH2F alkenes; and (6) miscellaneous, e.g., hydromonofluoromethylation of activated alkenes with ICH2F, or reductive desymmetrization of α-CH2F-substituted malonic ester.
Despite significant progress, however, to our knowledge, there is a lack of comprehensive review articles to summarize the advances in the catalytic enantioselective construction of CH2F-substituted stereocenters13; in sharp contrast, many elegant reviews have been published to introduce advances in the selective difluoro- or trifluoroalkylation.6 In light of this, we feel it is necessary to present a timely review that focuses on elucidating the evolution of constructing CH2F-substituted stereocenters, outlining the synthetic opportunities still open, thus providing related researchers with some reference and inspiration to develop more useful and diverse approaches for installing chiral organofluorine compounds bearing CH2F-substituted stereocenters. In the following section, we introduce advances according to the classification shown in Fig. 2.
In 2020, Gouverneur, Pupo and coworkers accomplished the first enantioselective ring opening of azetidinium triflates with fluorides by hydrogen bonding phase-transfer catalysis, which allowed the facile synthesis of enantioenriched γ-fluoroamines featuring an α-CH2F substituted stereocenter (Scheme 2).18 Using a 5–10 mol% BINAM-derived N-isopropyl-bis(urea) catalyst 14, a series of azetidinium triflates 13 worked effectively with CsF to afford optically active γ-fluoroamines 15 in 51–99% yields and 65–94% ee. It was noted that chiral γ-fluoroamines bearing a tetrasubstituted stereocenter were also obtained from the corresponding disubstituted azetidinium triflates, albeit with moderate enantioselectivity. In addition, it was found that the N-substituents of the azetidinium triflates influenced the reactivity, but its configuration was not important for enantioselectivity. A possible transition state was proposed in which a chiral N-alkylated bis-urea catalyst served as the hydrogen bond donor to bring CsF in the solution, and the generated chiral urea–fluoride complex ion pairing interacted with aziridinium ions. Enantioselective nucleophilic fluorination was used to deliver the targets. Furthermore, the use of the current method was highlighted by the efficient synthesis of fluorinated analog 17 of lorcaserin, a selective serotonin 2C receptor agonist approved by the FDA for chronic weight management.
In 2018, during their study in exploring Pd-catalyzed asymmetric allenylation of bis(phenylsulfonyl)methane or diethyl malonate with racemic 2,3-allenyl acetates, the Ma group investigated the asymmetric monofluoromethylation of racemic 2,3-allenyl acetate 27 with FBSM. The desired fluorobis(phenylsulfonyl)methylated allenylic product 29 was isolated in 81% yield with 97% ee using [Pd(π-cinnamyl)Cl]2 (2.5 mol%), (R)-DTBM-Segphos 40 (6 mol%), and K2CO3 (2.0 equiv.) (Scheme 3).21 A monofluoromethylated chiral allene 30 bearing both axial and central chirality was readily obtained in 96% ee after reductive desulfonylation with Mg/MeOH.
Later in 2009, You, Zhao, and coworkers reported regio- and enantioselective allylic alkylation of FBSM with allylic carbonates 31 catalyzed by a chiral iridium complex consisting of chiral phosphoramidite 32 (4 mol%) and [Ir(COD)Cl]2 (2 mol%), which allowed the access of a variety of α-fluorobis(phenylsulfonyl)methylated chiral allylic compounds 33 with 28–96% yields, 86:14 to >99:1 b/l and 70–96% ee (Scheme 4).22 The practicability of the current method was demonstrated by the gram-scale synthesis of 33g and its elaboration to monofluoromethylated (R)-ibuprofen in two steps without erosion of the enantioselectivity. In addition, monofluoromethylated (S)-ibuprofen and (R)- or (S)-naproxen were obtained from the corresponding fluorobis(phenylsulfonyl)methylated chiral allylic products.
With their research interest in organofluorine chemistry and asymmetric catalysis,23 the Yu group in 2022 established the highly regio- and enantioselective hydromonofluoromethylation of 1,3-dienes with FBSM 18 using low-cost and redox-neutral chiral nickel catalysis (Scheme 5).24 Under the catalysis of P-chiral (S,S)-QuinoxP* 36 (5.5–11 mol%) and Ni(COD)2 (5–10 mol%), various aromatic and aliphatic 1,3-dienes 35 reacted well to deliver fluorobis(phenylsulfonyl)methylated chiral allylic molecules 37 in 48–99% yields and 90–99% ee. Interestingly, they further realized a tandem hydromonofluorobis-(arylsulfonyl)methylation/reductive desulfonylation sequence, which allowed the access of enantioenriched α-CH2F or α-CD2F- substituted allylic molecules 38 with excellent ee. This method not only provides a facile strategy to construct diverse monofluoromethyl- or monofluoroalkyl-containing chiral allylic compounds but also represents the first catalytic enantio- and regioselective Markovnikov hydrofluoroalkylation of alkenes.
One year later, the Yu group utilized chiral nickel catalysis to accomplish the first asymmetric hydromonofluoromethylation of 1,3-enynes with FBSM (Scheme 6).25a A combination of chiral (R)-MeO-BIPHEP 40 (11 mol%) and Ni(COD)2 (10 mol%) effectively mediated the reaction and afforded a variety of monofluoromethyl-tethered axially chiral allenes 41 in 39–98% yields and 53–94% ee, in the presence of DABCO (1.0 equiv.). Notably, this constitutes the first asymmetric 1,4-hydrofunctionalization of 1,3-enynes using low-cost chiral nickel catalysis and opens an avenue for the activation of 1,3-enynes in developing asymmetric reactions. The mild reaction conditions, excellent substrate scope, and functionality tolerance highlighted the usefulness of this method. Moreover, tandem hydromonofluorobis(arylsulfonyl)methylation of a 1,3-enynes/reductive desulfonylation sequence was successfully achieved, which enabled the unprecedented preparation of CH2F- or CD2F-tethered chiral allenes 42 with 85–91% ee, which are otherwise inaccessible. In addition, Yu and coworkers explored the asymmetric hydromonofluoromethylation of alkoxyallenes, which delivered chiral monofluoromethylated allylic ethers with 68% ee when an (R)-tBu-PHOX-decorated chiral Pd complex was used as the catalyst.25b
Aside from chiral metal-catalyzed monofluoromethylation using FBSM, asymmetric organocatalysis has also demonstrated its potential in developing enantioselective transformations involving FBSM. In 2007, using 5 mol% of N-benzylquinidinium chloride 44a as the chiral phase-transfer catalyst, Shibata, Toru, and coworkers developed an enantioselective monofluoromethylation of aryl or alkyl α-amido sulfones 43 with FBSM, and a variety of enantiomerically enriched N-Boc α-fluorobisphenylsulfonyl amines 45 were obtained in 70–98% yields with 87–99% ee (eqn (1), Scheme 7).26 Later, in 2013, the Shibata group further employed chiral PTC catalysis to realize an asymmetric monofluormethylation of 2-aryl-3-(1-arylsulfonylmethyl)indoles 46 with FBSM. The use of 10 mol% cinchonine-derived PTC catalyst 47 with 1.2 equiv. Cs2CO3 afforded various monofluoromethyl substituted C2-arylindoles 48 with up to 97% ee (eqn (2), Scheme 7).27 The in situ generated vinylogous imines from 2-aryl-3-(1-arylsulfonylmethyl)indoles were the real active electrophiles, which reacted with FBSM to deliver the desired chiral adducts using the chiral catalyst. Moreover, both N-Boc α-fluorobisphenylsulfonyl amines 45 and fluorobisphenylsulfonyl-substituted C2-arylindole 48a readily underwent reductive desulfonylation with Mg/MeOH to give their corresponding monofluoromethylated derivatives.
Scheme 7 Chiral PTC -catalyzed enantioselective monofluoromethylation of α-amido sulfones or 2-aryl-3-(1-arylsulfonylmethyl)indoles with FBSM. |
Also, by chiral PTC catalysis, Shibata and coworkers achieved the highly enantioselective Michael addition of FBSM to α,β-unsaturated ketones (Scheme 8).28 A series of β-alkyl or β-aryl-substituted α,β-unsaturated ketones 51 worked well with FBSM under the catalysis of 5 mol% quinidine-derived quaternary ammonium salt 44b and 3.0 equiv. Cs2CO3 to afford β-fluorobis(phenylsulfonyl)methylated ketones 52 in 32–91% yields with 85–98% ee. The elaboration of product 52a into monofluoromethylated derivatives 53 and 54 further demonstrated the practicability of this protocol.
As shown in Scheme 8, a possible transition state was proposed, in which the hydroxyl group of catalyst 44b captured enones via hydrogen-bonding interaction with the carbonyl group, while the aromatic π–π interactions between enones and the catalyst might stabilize the transition state. Finally, the FBSM approached the enone from its Re-face because the Si-face was effectively blocked by the bulky parts of the benzyl substituent in the catalyst.
One year later, Córdova and coworkers reported the highly enantioselective Michael addition of FBSM to α,β-unsaturated aldehydes using asymmetric iminium catalysis (Scheme 9).29 It was found that the use of 20 mol% TMS-protected diarylprolinol 56 and 10 mol% Et3N proved to be the optimal catalyst system, and various β-alkyl or phenyl substituted enals 55 reacted smoothly with FBSM to afford the desired γ-fluorobis(phenylsulfonyl)methyl alcohols 57 in 45–84% yields and 84–95% ee after in situ reduction with NaBH4. A possible reaction involving chiral iminium catalysis was proposed, in which chiral secondary amine activated enals 55 to form the active iminium intermediate I, which reacted with FBSM to give the target, as shown in Scheme 9. Moreover, the monofluoromethylated alcohol could be readily obtained from product 57via reductive desulfonylation with Mg/MeOH.
In 2011, Tan, Jiang, Huang and coworkers established the highly enantioselective allylic alkylation of MBH carbonates 59 with FBSM 18 catalyzed by cinchona alkaloid-derived chiral tertiary amine (eqn (1), Scheme 10).30 In the presence of 10 mol% C2-symmetric (bis)cinchona alkaloid derivative (DHQD)2AQN 58, the desired products 60 were obtained in 58–72% yields with >99.9% ee for all cases. It should be noted that the synthetic usefulness of this method was highlighted by the diastereoselective synthesis of enantiopure β-methyl-γ-monofluoromethyl alcohol 62 from the monofluoromethylated allylic alkylation adduct 60a in three steps. Almost simultaneously, the Shibata group also reported the allylic alkylation of MBH carbonates with FBSM using 10 mol% (DHQD)2AQN, and they further improved the enantioselectivity of this reaction using a cooperative catalyst system consisting of (DHQD)2AQN with Lewis acid FeCl2 or Ti(OiPr)4 (eqn (2), Scheme 10).31 In the presence of nBu3SnH and AIBN, an intramolecular radical cyclization of 60b proceeded smoothly to afford dihydroindene derivative 63, which could be converted into 1-monofluoromethylindene 64.
In 2015, Hamashima et al. established the highly enantioselective fluorolactonization of o-vinylbenzoic acids 69 with electrophilic Selectfluor 65 using a newly designed bifunctional hydroxyl carboxylate catalyst 70 (Scheme 12).33 The corresponding monofluoromethylated isobenzofuranones 71 were obtained in 40–99% yields and 44–94% ee values, in the presence of 10 mol% axially chiral bifunctional hydroxyl carboxylic acid 70a and 1.5 equiv. Na3PO4. Control experiments revealed that the presence of a primary hydroxyl group at an appropriate position in the catalyst is important to achieve high asymmetric induction because the use of catalyst 70b afforded the target with 68% ee, while almost no enantioselectivity was observed when using precatalysts 70c–e under otherwise identical conditions. As shown in Scheme 12, a plausible mechanism was proposed in which catalyst 70a was first deprotonated by potassium phosphate and tended to aggregate, and a binary complex I consisting of the catalyst and substrate was then formed after dissociation enabled by the substrate. Subsequently, complex I was reacted with Selectfluor 65 to generate active fluorinated species II in the liquid phase. Finally, targeted fluorolactones 71 were obtained after fluorolactonization, accompanied by the formation of species III. It should be noted that the hydroxyl group within the catalyst might play a crucial role in the interaction with o-vinylbenzoic acids by forming a hydrogen bond with the neighboring carboxylate anion to fix its position.
Four years later, the Hamashima group reported the enantioselective 5-exo-fluorocyclization of ene-oximes 72 with Selectfluor 65 using an axially chiral carboxylate phase transfer catalysis (Scheme 13).34 A combination of 10 mol% linked-binaphthyl dicarboxylic acid precatalyst 7335 with 1.5 equiv. Na3PO4 furnished the desired monofluoromethylated isoxazolines 74 in up to 69% yield and 84% ee. The aliphatic ene-oxime was less reactive for fluorocyclization because only a 12% yield of 74d was observed in the case of cyclohexyl-substituted ene-oxime even at room temperature. Notably, the hydrogen-bonding interaction between the hydroxyl in ene-oxime and the dianionic catalyst was essential for this transformation, as no reaction occurred when O-methylated ene-oxime 72f was subjected to the standard reaction conditions.
In 2017, Hoveyda and coworkers developed an efficient method for the enantioselective construction of monofluoromethylated Z-homoallylic tertiary alcohols through chiral boron-catalyzed asymmetric allylation of α-CH2F aryl ketones 75 (Scheme 15).39 By using a combination of aminophenol 79 (2.5 mol%) with Zn(OMe)2 (5 mol%), both Z- and E-γ-substituted boronic acid pinacol esters 78 performed well with several α-CH2F aryl ketones, delivering the corresponding α-addition products 80 in up to 94% yield, 94% ee, >98:2 α:γ selectivity with >78:22 Z:E selectivity, and their diastereomers 80’ in up to 98% yield, 92% ee, >98:2 α:γ and Z:E selectivity. Mechanistic studies revealed that the presence of the fluoroalkyl group was the key factor for the formation of the higher energy α-addition Z-selective products, which might be ascribed to the electrostatic and steric interactions between the catalyst and monofluoromethyl group, as shown in the proposed transition state of Scheme 15. In addition, the presence of the Lewis acidic Zn(OMe)2 might enhance the activity and efficiency of α-CH2F aryl ketones 75 by reducing the carbonyl-boron coordination.
β-CH2F-Substituted α,β-unsaturated carbonyl compounds have also found applications in forging CH2F-substituted stereocenters. In 2017, Schaus, Thomson and coworkers investigated the enantioselective reductive allylation of β-CH2F-β-phenyl disubstituted enal 87 during the study of the asymmetric traceless Petasis borono-Mannich reaction of various enals with 2-nitro-benzenesulfonyl hydrazide and allylboronate (Scheme 17).42 The in situ generated α,β-unsaturated hydrazone, from enal 87 and 2-nitro-benzenesulfonyl hydrazide 89, reacted smoothly with allyl boronate 88 to give the desired 1,4-diene 91 bearing a CH2F-substituted stereocenter with 77% yield and 86% ee after undergoing allylic diazene rearrangement.
In 2018, the Zhang group reported the ligand-controlled Cu(I)-catalyzed asymmetric 1,3-dipolar cycloaddition of α-substituted iminoester 93 with β-CH2F-β-aryl disubstituted enone 92, which provided an efficient method for the enantioselective regiodivergent synthesis of CH2F-substituted chiral pyrrolidine bearing two adjacent quaternary stereocenters or two discrete quaternary stereocenters (Scheme 18).43 A combination of Cu(CH3CN)4BF4 (10 mol%) with P,N-type ligand (S,Sp)-iPr-Phosferrox 94 (11 mol%) afforded monofluoromethylated pyrrolidine 95 bearing two adjacent quaternary stereocenters with 65% yield, 6:1 rr and 99% ee, while pyrrolidine 97 with two discrete quaternary stereocenters was obtained in 71% yield, 11:1 rr, and 94% ee when using (S,Rp)-PPFA 96 as a chiral ligand. In the same year, they employed their Ming-Phos 99 derived chiral copper complex to realize the asymmetric 1,3-dipolar cycloaddition of glycine ketimines 98 with β-CH2F β-aryl disubstituted enone 92, delivering product 100, featuring a CH2F substituted stereocenter with 61% yield, 6:1 dr and 94% ee (Scheme 18).44
Asymmetric cyclopropanation of diazo compounds with CH2F substituted alkenes by either chiral rhodium, gold catalysis or biocatalysis is identified as a powerful strategy for the stereoselective preparation of monofluoromethylated cyclopropanes.45 In 2018, Zhou, Ma and coworkers developed the highly diastereo- and enantioselective cyclopropanation of unprotected diazooxindoles 101 with α-CH2F styrenes 102 catalyzed by an SKP-derived chiral digold complex, allowing access to optically active spirocyclopropyloxindoles 104 with two adjacent quaternary stereocenters featuring a CH2F group (Scheme 19).46 The use of cationic digold complex derived from 4.0 mol% 103 (AuCl)2 and 4.4 mol% AgOTf delivered a series of CH2F-substituted cyclopropanes 104 in 57–93% yields, >20:1 dr and 87–95% ee. Interestingly, control experiments revealed a dramatic rate of acceleration when fluorobenzene was used as the solvent. By theoretical calculation study, the authors proposed that the formation of strong C–F⋯H–N interactions47 between the solvent fluorobenzene and the N–H bond of diazooxindole-derived Au(I)–carbenoid intermediate played a key role, as shown in the proposed reaction mechanism in Scheme 19.
Scheme 19 Au-catalyzed asymmetric cyclopropanation of unprotected diazooxindoles 101 with α-CH2F styrenes. |
Meanwhile, the same group found that the use of 0.5 mol% Rh2(R-DOSP)4 could mediate the asymmetric cyclopropanation of aryl diazoacetates 105 with α-CH2F styrenes 102, and afforded CH2F-substituted chiral cyclopropanes 106 with adjacent quaternary stereocenters in 66–79% yields, >20:1 dr and 91–95% ee values (eqn (1), Scheme 20).46 Shortly after, Jubault and coworkers also realized the same cyclopropanations by using 1 mol% of Rh2(S-BTPCP)4, providing chiral cyclopropanes 106 with 24–99% yields, >20:1 dr and 55–98% enantioselectivities (eqn (2), Scheme 20).48 Later, in 2022, the Zhou group accomplished the highly diastereo- and enantioselective cyclopropanation of diazothiooxindoles 107 with α-CH2F styrenes catalyzed by 0.5 mol% of Rh2(S-TCPTTL)4 (eqn (3), Scheme 20).49 Several chiral CH2F substituted spirocyclopropylthiooxindoles 108 bearing adjacent quaternary stereocenters were obtained in up to 96% yield, 18:1 dr, and 97% ee. By using 1 mol% of Rh2(S-BTPCP)4 as the catalyst, Zhu, Wu, and coworkers recently reported an asymmetric intramolecular cycloisomerization of CH2F substituted enyne for the enantioselective synthesis of functionalized bicycle featuring a CH2F substituted stereocenter.50 In addition to chiral rhodium catalysis, Fasan, Jubault and coworkers attempted the enantioselective cyclopropanation of ethyl 2-diazoacetate 105a with α-CH2F styrene 102a using Mb(H64V, V68A) as a biocatalyst, which delivered the desired CH2F substituted cyclopropane 106a with 99% yield, 99% de, and 98% ee (eqn (4), Scheme 20).51
Scheme 20 Asymmetric cyclopropanation of α-CH2F styrenes catalyzed by chiral Rh complexes or enzymes. |
Aside from asymmetric cyclopropanation, the asymmetric dihydroxylation of α-CH2F styrenes 102 was also explored by Jubault, Poisson, and coworkers in 2019 (Scheme 21).52 Under the catalysis of AD-mix-α or AD-mix-β (Os: 0.2 mol%), several optically active α-monofluoromethylated tertiary alcohols 109 were obtained in up to 93% yield and 98% ee. Additionally, various α-difluoromethylated styrenes were also tolerated, giving rise to the corresponding α-difluoromethylated chiral tertiary alcohols with excellent enantioselectivities.
In 2022, Andersson, Zhou and coworkers developed an efficient protocol for enantioselective synthesis of monofluoromethylated stereocenters by chiral Ir-catalyzed enantioselective hydrogenation of fluorinated alkenes (Scheme 22).53 It was found that (E)-ethyl 4-fluoro-3-phenylbut-2-enoate 110 was efficiently hydrogenated with 5 bar H2 to give ethyl (R)-4-fluoro-3-phenylbutanoate 112 featuring a CH2F stereocenter with 85% ee, in the presence of 0.5 mol% Ir-N,P catalyst 111a. Moreover, the use of 1 mol% Ir-N,P catalyst 111b enabled the hydrogenation of various vinyl fluorides 113 to furnish a series of monofluoromethylated chiral molecules 114 in up to 99% yield and 98% ee values.
Scheme 23 Organocatalyzed desymmetrization of glutaric anhydrides and photoredox-catalyzed decarboxylic fluorination. |
As shown in Scheme 23, a plausible reaction mechanism for dicarboxylic fluorination was proposed, in which the photoexcitation of the ground-state photocatalyst Ir(III) by visible light resulted in excited *Ir(III), which then underwent oxidative quenching in the presence of Selectfluor to give Ir(IV). Subsequently, the resultant chiral glutaric acid monoester 118 underwent deprotonation, oxidation, and sequential elimination of CO2 with the assistance of both the Na2HPO4 and oxidized photocatalyst Ir(IV) to afford the alkyl radical intermediate and re-generate photocatalyst Ir(III). Finally, chiral fluorides 120 were generated after fluorine transfer from Selectfluor to the alkyl radical species.
In the same year, the Huang group reported a chiral Zn-catalyzed asymmetric hydrosilylation of α-CH2F substituted malonic ester 121, which allowed the access of a CH2F substituted stereocenter (Scheme 24).55 They found that chiral α-CH2F substituted β-hydroxylester 123 was selectively obtained in 74% yield with 68% ee in the presence of 10 mol% of newly developed tetradentate chiral ligand 122a, 20 mol% of diethylzinc, and 3.0 equiv. of (MeO)3SiH.55a Interestingly, the use of a higher concentration of (MeO)3SiH (6.0 equiv.) and pipecolinol-derived tetradentate ligands 122b instead of 122a enabled the aldehyde-selective desymmetrization, and afforded α-CH2F substituted aldehyde 124 in 71% yield and 96% ee, with 6.8:1 aldehyde/alcohol selectivity.55b
Recently, Li, Huang, Xia and coworkers developed a nickel-catalyzed enantioselective hydromonofluoromethylation of enamides or enol esters 125 with ICH2F 126 as a monofuoromethyl reagent (Scheme 25).56 Combining 12 mol% cyclopropanyl substituted chiral Box ligand 127 and 10 mol% NiI2 enabled the construction of various chiral α-monofluoromethylated amides and esters 128 in up to 96% yield and 99% ee in the presence of KF and poly(methylhydrosiloxane) (PMHS). Moreover, the practicability of this method was demonstrated by the preparation of chiral oxazoline 129 (42% yield, 94% ee) from 128fvia reduction with LiAlH4, as well as the synthesis of α-monofluoromethylated tetrahydroquinoline 130 (70% yield, 84% ee) from 128g via Buchwald–Hartwig amination.
Footnote |
† These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2024 |