Jiaqiong
Sun
b,
Guangfan
Zheng
a,
Ge
Zhang
a,
Yan
Li
*a and
Qian
Zhang
*ac
aJilin Province Key Laboratory of Organic Functional Molecular Design & Synthesis, Department of Chemistry Northeast Normal University, Changchun 130024, China. E-mail: liy078@nenu.edu.cn; zhangq651@nenu.edu.cn
bSchool of Environment, Northeast Normal University, Changchun 130117, China
cState Key Laboratory of Organometallic Chemistry, Shanghai Institute of Organic Chemistry, Chinese Academy of Sciences, 345 Lingling Lu, Shanghai 200032, China
First published on 14th January 2025
This review offers a comprehensive overview of recent advancements in the asymmetric construction of C–N bonds involving radical intermediates. Enantioselective radical amination strategies have proven to be highly effective for synthesizing chiral amines and nitrogen-containing heterocycles. Significant progress has been made in the enantioselective installation of N-containing groups into halogenated alkanes, olefins, and dienes with the asymmetric formation of C–N bonds as the key step via diverse pathways including reductive elimination, radical-polar crossover, amino group substitution, radical–radical cross-coupling, etc. This review highlights these recent developments and the mechanistic insights that drive these transformations.
Significant progress has been made in enantioselective radical amination (Scheme 1). Mechanistically, the first in situ generated sp2-hybridized alkyl radicals serve as critical intermediates. These alkyl radicals can be formed through various methods, such as single electron transfer (SET) or halogen atom transfer (XAT) of alkyl halides, hydrogen atom transfer (HAT) from alkanes, radical addition to olefins and so on. Depending on the enantio-determining step, the subsequent C–N formation strategies can be categorized into four main classes: (1) reductive elimination of chiral alkyl-metal complexes formed by the interaction between alkyl radicals and transition metals with chiral ligands; (2) radical-polar crossover: a chiral catalyst-bound nitrogen source coupled with a carbocation or hypervalent chiral alkyl-metal species undergoing SN2-like nucleophilic amination; (3) radical substitution of alkyl radicals with chiral N-metal species; and (4) radical–radical cross-coupling of N-centered radicals and alkyl radicals. Additionally, sporadic enantioselective addition of N-centered radicals to olefins has also been reported. As a result, a wide range of substrates, including halogenated hydrocarbons, alkanes and alkenes, could be transformed into value-added chiral amines or N-containing heterocycles via these enantioselective radical amination strategies.
![]() | ||
Scheme 1 Representative strategies for enantioselective construction of C–N bonds involving radical intermediates. |
This review offers an overview of the latest advancements in the asymmetric formation of C–N bonds that involve radical intermediates. Enzyme-facilitated strategies for C–N bond formation are not covered.19,20 By highlighting the fundamental principles, strategies, and key intermediates involved in these transformations, this review aims to inspire further research and innovation in the field of asymmetric radical amination chemistry.
In 2022, the same group26 provided detailed mechanistic insights into the key enantioselective C–N bond formation step of their previous report, a visible light-induced Cu-catalyzed enantioconvergent amination of carbazole and tertiary alkyl halides with an amide carbonyl group at the α-position.27 They synthesized and characterized a CuII model complex, (NP)CuIIcab′ I, which could couple with trityl radicals in toluene at −40 °C to produce N-alkylated carbazole in an excellent yield (91%). EPR studies and the corresponding DFT calculations supported the idea that the CuII–N complex functioned as a CuII–N metalloradical during the catalyzed coupling process. The authors proposed a possible mechanism as follows: L*CuICl reacts with nitrogen nucleophiles to generate the L*CuIcarb complex II, which then undergoes irradiation to produce an excited state, [L*CuIcarb]* III. III subsequently participates in a SET with alkyl halides, resulting in the formation of an alkyl radical and the [L*CuIIcarb]Cl complex IV, which reacts with another nitrogen nucleophile to form V. The formation of C–N bonds through VI yields the desired products with good enantioselectivities (Scheme 3).
![]() | ||
Scheme 3 Investigation of the C–N bond-forming step in a photoinduced, Cu-catalyzed enantioconvergent N-alkylation. |
In the same year, Fu and colleagues28 developed a photoinduced, copper-catalyzed enantioconvergent C–N coupling reaction of racemic alkyl halides with aniline derivatives. This method efficiently synthesizes chiral amines featuring a fully substituted stereocenter. A variety of racemic tertiary α-substituted α-chloro-/bromo-nitriles were identified as suitable electrophiles for producing enantioconvergent N-alkylation products with good enantioselectivities. However, the yield of the reaction was influenced by the size of the substituent at the α-position. Aniline derivatives with a substituent at the para-position yielded products in moderate to good yields and demonstrated good enantioselectivities. In contrast, when aniline was used as the nucleophile, a substantial amount of electrophilic addition occurred at the para position. Mechanistically, irradiation of P*2CuICl I under blue LEDs generated the excited-state intermediate II. This intermediate underwent SET with alkyl halides to form CuII-species III and an alkyl radical. The coordination of aniline with III produced the intermediate IV, which then combined with the alkyl radical to yield the desired product with good enantioselectivity, simultaneously regenerating I (Scheme 4).
![]() | ||
Scheme 4 Photoinduced, Cu-catalyzed enantioconvergent alkylations of anilines by racemic tertiary electrophiles. |
In 2021, the Liu group29 developed a copper-catalyzed enantioconvergent radical C–N coupling method using diverse racemic secondary alkyl halides, employing sulfoximines as effective ammonia surrogates. This reaction efficiently produces highly enantioenriched N-alkyl sulfoximines featuring secondary benzyl, propargyl, α-carbonyl alkyl, and α-cyano alkyl stereocenters. The employment of electron-rich multidentate anionic N,N,P-ligands significantly enhances the reducing capability of CuI catalysts, allowing for the efficient generation of alkyl radicals from alkyl halides under mild thermal conditions, facilitating the enantioconvergent radical C–N coupling of racemic secondary alkyl chlorides and bromides. Mechanistically, the CuI-sulfoximine complex I undergoes SET with the racemic alkyl halides, generating a secondary alkyl radical II and a CuII-complex III. Subsequent enantio-determining C–N bond formation might occur, yielding highly enantioenriched N-alkyl sulfoximine derivatives (Scheme 5).
In 2023, the Liu group30 designed and synthesized a new anionic N,N,N-ligand with a long spreading side arm. By means of this ligand, they extended the copper-catalyzed enantioconvergent radical C(sp3)–N coupling from secondary alkyl halide electrophiles to spatially crowded tertiary alkyl halides. Upon using sulfoximines as the nitrogen sources, a series of α-aminocarbonyl tertiary alkyl chlorides smoothly converted into α,α-disubstituted amino acids in good yields with high enantioselectivities (Scheme 6).
![]() | ||
Scheme 6 Cu-catalyzed enantioconvergent radical C(sp3)–N cross-coupling for access to α,α-disubstituted amino acids. |
In 2023, Liu and co-workers31 reported the utilization of chiral tridentate anionic ligands to facilitate the Cu-catalyzed enantioconvergent N-alkylation of aliphatic amines and α-carbonyl alkyl chlorides. This method could directly convert bulk feedstock chemicals, such as ammonia, methylamine, dimethylamine and pharmaceutically relevant amines, into chiral α-amino amides. This reaction displayed excellent enantioselectivity and functional group compatibility. Furthermore, the robust method was successfully applied in the late-stage functionalization and expedited synthesis of various amine drug molecules. The key to the achievement could be attributed to the chiral tridentate anionic N,N,N-ligands, which exhibit strong binding affinities to metal catalysts, not only overcoming catalyst poisoning by aliphatic amines or ammonia but also inducing high enantioselectivities. The authors proposed a possible mechanism for this reaction. It begins with the formation of CuI species I, which undergoes intramolecular oxidative addition to produce species II and III, which are in equilibrium. The subsequent outer-sphere attack of III by an amine leads to the enantioselective formation of intermediate IV. Finally, a ligand exchange with an alkyl chloride releases the chiral aliphatic amines and regenerates intermediate I, closing the catalytic cycle (Scheme 7). In 2024,32 the same group expanded the range of nucleophiles to include bulky secondary and primary ones. They developed an outer-sphere nucleophilic attack mechanism to circumvent the difficulties of transmetalation that arise with sterically congested nucleophiles. This approach offers new opportunities for the construction of chiral carbon centers, particularly challenging sterically congested ones.
At the same time, Fu and colleagues33 independently developed a copper/chiral isoxazoline-catalyzed enantioconvergent substitution of racemic α-chloro/bromo-N-phenylbutanamide using amines, leading to direct access to chiral α-amino amide derivatives in good to excellent yields with excellent enantioselectivities. Both aromatic and alkyl amines were compatible with this radical amination process (Scheme 8).
In 2018, Zhang37 and coworkers developed a groundbreaking CoII-catalyzed enantioselective amination of C(sp3)–H bonds employing CoII-based MRC. This method involved the activation of sulfamoyl azides by the Co-MRC system, which generated α-CoIII-aminyl radicals I. These radicals facilitated a 1,6-HAT followed by enantioselective C–N bond formation, producing six-membered chiral heterocyclic sulfamides 27 with high yields and excellent enantioselectivities. The process displayed broad substrate compatibility, including benzylic, allylic, and propargylic C(sp3)–H bonds. Moreover, chiral cyclic sulfamides could be readily converted into valuable 1,3-diamines without loss of enantiopurity (Scheme 11, top).
Furthermore, the CoII-based MRC-catalyzed 1,6-C(sp3)–H amination system was further extended to the racemic tertiary C(sp3)–H bonds.38 As shown in Scheme 11 (bottom), this approach displayed a broad substrate scope, accommodating various ester and aryl functionalities, and enabled the efficient synthesis of chiral six-membered cyclic sulfamides bearing quaternary stereocenters 28. Detailed mechanistic studies, including KIE analysis and EPR detection of intermediates, provided valuable insights into the enantioselective control of the radical process. In situ generated α-CoIII-aminyl radicals underwent 1,6-HAT of tertiary C–H bonds, resulting in the prochiral tertiary alkyl radical center III, which then underwent enantio-determining radical substitution, leading to the construction of quaternary stereogenic centers. The stepwise radical mechanism, featuring HAT and radical substitution, was key in achieving enantio-convergent amination of tertiary C–H bonds.
In 2019,39 the same group developed an enantioselective CoII-catalyzed radical 1,5-C–H amination of sulfonyl azides 29. Using Co-2 (for arylsulfonyl azides) or Co-3 (for alkylsulfonyl azides), both types of azides underwent efficient C–H amination, producing chiral cyclic sulfonamides 30 with high yields and excellent enantioselectivities (Scheme 12, top). The reaction effectively accommodated a wide variety of functional groups, including benzylic, allylic, and non-activated C(sp3)–H bonds. Mechanistically, the activation of organic azides by CoII-MRC leads to the formation of α-CoIII-aminyl radicals I. These radicals undergo a 1,5-HAT, followed by an enantio-determining amino-radical substitution, which enables the enantioselective formation of five-membered cyclic sulfonamides. The involvement of α-CoIII-aminyl radical I was confirmed through isotropic EPR spectroscopy.
Later, they developed a novel cobalt(II)-catalyzed enantio-divergent radical 1,5-C(sp3)–H amination of sulfamoyl azides, employing D2-symmetric chiral porphyrins as ligands.40 By modulating the cavity environment of the ligands (Co-4 or Co-6), they allowed the enantio-divergent formation of the product. Detailed deuterium-labelling studies and DFT calculations revealed a unique mechanism of asymmetric induction, combining enantio-differentiative HAT with stereo-retentive radical substitution (Scheme 12, bottom).
In 2022, the Zhang41 group further extended the CoII-based MRC system by combining radical and ionic approaches for the enantioselective synthesis of β-functionalized chiral amines from alcohols (Scheme 13). This innovative strategy merges an enantioselective radical process, 1,5-C–H amination of alkoxysulfonyl azides, with an enantiospecific ionic process for the ring-opening of the resulting five-membered cyclic sulfamidates. This method provides an efficient and practical route to synthesize highly enantioenriched β-functionalized amines with a broad substrate scope, high yields, and excellent stereoselectivities. Mechanistically, the CoII-MRC activates alkoxysulfonyl azides to generate α-CoIII-aminyl radicals, which undergo 1,5-HAT and radical substitution to form chiral cyclic sulfamidates 36. These intermediates can then undergo nucleophilic ring-opening, allowing access to a wide range of chiral amines 37. This approach is compatible with various C–H bonds and nucleophiles, significantly expanding the synthetic toolbox for chiral amine construction.
![]() | ||
Scheme 13 CoII-based MRC catalysis combining radical and ionic approaches for the enantioselective synthesis of β-functionalized chiral amines. |
The Co-MRC system could be extended to the enantioselective amination of simple benzylic and allylic C–H bonds through the intermolecular HAT process, providing new opportunities for the efficient and precise construction of chiral nitrogen-containing compounds.
In 2020, the Zhang42 group reported an enantioselective intermolecular radical C–H amination of benzylic C–H bonds bearing carboxylic acid esters by employing fluoroaryl azides as the nitrogen sources, which facilitated the construction of valuable chiral α-amino acid derivatives (Scheme 14, top). The key to their success was the development of D2-symmetric chiral amidoporphyrin ligands, which enhanced noncovalent interactions and controlled both reactivity and enantioselectivity. The intermolecular C–H amination was conducted under mild conditions and demonstrated a broad substrate scope, yielding chiral amines with high chemo-selectivity and excellent enantioselectivity (40). Mechanistic studies uncovered a stepwise radical pathway that involved metalloradical activation of organic azides, HAT from C–H substrates, and enantio-determining radical substitution, leading to C–N bond formation.
The convergent transformation of various isomeric mixtures of alkenes poses a significant challenge in organic synthesis chemistry, requiring innovative solutions. In 2023, Zhang43 and colleagues introduced a novel approach for the intermolecular asymmetric radical allylic C–H amination of these isomeric mixtures (Scheme 14, bottom). By utilizing CoII-based MRC and modularly designed D2-symmetric chiral amidoporphyrin ligands, this research presents a highly selective approach for synthesizing chiral α-tertiary amines 42 from organic azides and isomeric mixtures of alkenes. The versatility of this method is highlighted by its compatibility with a wide range of fluoroaryl azides and alkenes, enabling the efficient formation of allylic amines with excellent stereocontrol. Mechanistic studies, including DFT calculations and KIE analysis, reveal a stepwise radical mechanism and provide insights into the factors influencing regioselectivity and enantioselectivity. The key to achieving the chemical and stereochemical convergent transformation of isomeric mixtures of alkenes is the stepwise radical reaction mechanism and the essential property of allyl radical delocalization.
In 2020, Liu45 and colleagues developed a Cu/chiral phosphoric acid (CPA)-catalyzed asymmetric radical intramolecular 1,5-C–H amination of allylic and benzylic substrates. This method successfully produced enantioenriched α-alkenyl and α-arylpyrrolidines (Scheme 16). The process uses N-hydroxyphthalimide (NHPI) to mediate intermolecular HAT, generating alkyl radicals that facilitate enantioselective C–N bond formation. This approach demonstrates high efficiency and a broad substrate range, accommodating both electron-rich and electron-deficient substituents on the allylic and benzylic C–H bonds.
![]() | ||
Scheme 16 Cu/CPA-catalyzed intramolecular asymmetric 1,5-C–H amination of allylic and benzylic substrates. |
Later, the same research group46 reported an intramolecular, radical enantioconvergent amination of tertiary β-C(sp3)–H bonds of racemic ketones 56 catalyzed by Cu(I)/CPA (Scheme 17). The reaction demonstrates high efficiency and broad substrate compatibility, including aryl and heteroaryl ketones, with good to excellent enantioselectivity (up to 96% ee). Moreover, functional groups such as alkynes, phosphonates, and borates are well tolerated. Mechanistically, this method employs a CuI/CPA(L17) catalytic system that facilitates the generation of N-centred radicals I, which undergo intramolecular 1,5-HAT to form prochiral tertiary alkyl radicals II. II coordinated by CuI/CPA generates III. The subsequent intramolecular enantio-determining amino radical substitution of III delivers the desired products 59. In 2023, they further extended the scope of this system to N-chlorosulfonamide-type substrates for synthesizing pyrrolidine structural motifs featuring an α-quaternary stereocenter with high levels of enantiopurity.47
![]() | ||
Scheme 17 Cu/CPA-catalyzed intramolecular asymmetric 1,5-C–H amination of allylic and benzylic substrates. |
Independently, Nagib48 and coworkers developed a Cu/PC dual catalyzed intramolecular 1,5-C–H amination method for synthesizing chiral β-amino alcohols (Scheme 18). Mechanistically, addition of alcohol 60 to an imidoyl chloride 61 chaperone results in the formation of an oxime imidate 62, which selectively binds to a chiral Cu catalyst, generating complex I. Energy transfer between PC* and I leads to the generation of a Cu-bonded N-centred radical II. Regio- and enantioselective HAT followed by stereoselective C–N bond formation leads to the formation of chiral oxazoline 63. Subsequent hydrolysis yields the enantioenriched β-amino alcohol 64. Unlike Liu's research, this approach allows for chiral induction over two steps (both HAT and alkyl radical trapping), and the intermolecular C–N formation strategy is rapid enough to retain (and enhance) chiral memory. The method successfully delivers chiral β-amino alcohols from various alcohol substrates, including those containing alkyl, allyl, and benzyl groups, offering a versatile platform for synthesizing valuable chiral β-amino alcohols under mild conditions. In 2024,49 the same group further extended the enantioselective intramolecular C–H amination system to the synthesis of unprotected pyrrolines from oximes. The chiral pyrrolines can undergo reductions or nucleophilic additions with excellent stereocontrol, offering a modular platform for synthesizing enantioenriched chiral pyrrolidines.
In 2023, Chang50 and coworkers developed Cu/chiral BOX-catalyzed regio- and enantioselective δ-C(sp3)–H amidation of dioxazolones 65, enabling the construction of six-membered chiral lactams 66 with excellent regioselectivity and high enantioselectivities (>99:
1 er). Mechanistic studies reveal that the open-shell Cu-nitrenoid species play a critical role, mediating regioselective 1,6-HAT, followed by enantioselective radical rebound to form the C–N bond (Scheme 19).
In 2023, Zhou and colleagues51 developed a copper/chiral bisoxazoline (BOX)-catalyzed intermolecular enantioselective benzylic C(sp3)–H amination. This process uses a peroxide as both the oxidant and HAT reagent, with benzamide as the nitrogen source. Mechanistically, a tert-butoxy radical is generated in situ and initiates a rate-limiting HAT with the benzylic C(sp3)–H bond, forming a benzylic radical IV. This radical is then reversibly trapped by a CuII-amide complex III, leading to a Cu(III) intermediate V, which undergoes enantioselective reductive elimination to yield the chiral amide 68. This reaction accommodates a broad functional group range and has demonstrated its utility in synthesizing bioactive chiral amines, including (R)-Tecalcet 68a, Dapoxetine 68c, and Rivastigmine 68b, underscoring its potential for pharmaceutical applications (Scheme 20, top).
Independently, the Lian and Kramer group52 introduced a novel copper/PC dual-catalysis strategy for enantioselective benzylic C–H amination and amidation, using benzylic C–H substrates as the limiting reagent (Scheme 20, bottom). This method efficiently produces chiral benzylic amides and carbamate-protected benzylamines through direct C–H functionalization. It is scalable, compatible with various nitrogen sources, including cost-effective 15N-labeled reagents, and adaptable to drug synthesis.
![]() | ||
Scheme 22 Ruthenium-catalyzed enantioselective intramolecular C–H amination of urea derivatives or sulfamoyl azides. |
![]() | ||
Scheme 23 Enantioselective C–H amination catalyzed by nickel iminyl complexes supported by anionic bisoxazoline ligands. |
![]() | ||
Scheme 24 Chiral iron porphyrin-catalyzed enantioselective intramolecular C(sp3)–H bond amination upon visible-light irradiation. |
The enantioselective amination of readily available and abundant C–H bonds in alkanes represents a highly effective and straightforward strategy for synthesizing chiral amines. Radical methods can activate inert C–H bonds through the HAT process, generating highly reactive alkyl radicals that interact with transition metals to facilitate enantioselective C–N bond formation. This enantioselective radical C–H amination system offers compelling alternatives for the synthesis of chiral amines, amides, and azaheterocycles, which are vital in pharmaceuticals and organic synthesis.
Building on this research, the same group61 further expanded the utilization of this transformation by employing CuH to catalyze the asymmetric radical hydroamination of β-polyfluoroalkyl alkenes 91, a weak Michael receptor, thus facilitating the production of enantioenriched β-polyfluoroalkyl-substituted α-chiral tertiary alkylamines. The authors conducted density functional theory (DFT) calculations to shed light on the regio- and stereoselectivity of the reaction. As demonstrated in Scheme 27, bottom, the in situ-generated Cu–H species I undergoes migratory insertion into olefins to form a benzyl-CuI complex II. Subsequent SET between complex II and the electrophilic aminating reagent generates an alkylamine radical III, which is stabilized by chiral CuII species through coordination. The enantio-determining radical–radical coupling occurs between the alkyl radical V and the CuI-bounded dialkyl amine radical species VI, delivering the desired β-fluoroalkyl-substituted chiral amines and regenerating the Cu catalyst. Such an elegant asymmetric radical cross-coupling of in situ-generated nitrogen radicals with a carbon-based radical species to create an asymmetric C–N bond provides a promising new strategy for enantioselective C–N bond formation, making it a valuable tool in organic synthesis.
In 2020, Akai63 and coworkers reported an asymmetric hydroamination of nonactivated alkenes with benzotriazoles using an optically active Co(salen) complex, an N-fluoropyridinium salt, and a phenylsilane reagent. However, the enantioselective excess achieved was only up to 16%. When 5-phenyl tetrazole was utilized as the nitrogen source,64 the asymmetric hydroamination of 4-phenyl-1-butene did not exceed 20% ee with any of the tested chiral cobalt catalysts. The most favourable outcome was observed in the asymmetric hydroamination of indene with tetrazole, resulting in a 66% enantiomeric excess and 24% yield of chiral amine 94 (Scheme 29).
In 2021, the Zhang group65 successfully developed the first highly enantioselective hydroamination of alkenes utilizing N-fluorobenzenesulfonimides (NFSI) as both a nitrogen source and an oxidant under chiral cobalt catalysis (Scheme 30). This reaction features mild conditions, excellent regioselectivity, and wide functional group tolerance, and demonstrates a broad substrate scope. A variety of alkenes including styrenes, aliphatic alkenes and α,β-unsaturated carbonyl compounds can be engaged in this transformation to deliver the desired chiral amides with good to excellent yields. Mechanistic studies suggest that this reaction involves a cobalt(III)-hydride-mediated HAT to generate alkyl radicals III, which can be trapped by CoII species to generate the CoIII-alkyl intermediates IV. Notably, the CoIII-alkyl intermediates IV can be further oxidized to high-valent CoIV-alkyl intermediates, which act as electrophiles and are intercepted by a nitrogen source through an elegant SN2-substitution. Due to the involvement of radical intermediates, the enantiocontrol of the hydroamination of aliphatic alkenes is relatively lower than that of styrene. This reaction offers a novel and general method for enantioselective hydroamination.
![]() | ||
Scheme 30 Co-catalyzed asymmetric hydroamination of olefins employing NFSI as the oxidant and amination source. |
The catalytic asymmetric hydroamination of alkenes using Lewis basic amines has long been a challenging task in synthetic chemistry, due to the inherent strong coordination of Lewis basic amines that can potentially deactivate the chiral metal catalyst. Later in 2023, Zhang and colleagues developed a highly enantioselective hydroamination of aryl alkenes with secondary amines by leveraging cobalt hydride-mediated HAT, followed by a radical-polar crossover process involving high-valent cobalt(IV) species, significantly broadening the application of the approach (Scheme 31).66 Diarylamines, cyclic acyclic secondary anilines with various N-alkyl substituents, and dialkylamines are all valid nitrogen sources, producing the desired α-chiral tertiary amines with good to excellent yields and enantioselectivities. Some secondary amines bearing a –CN or pyridyl group, which easily coordinate with chiral metal centres, are also well accommodated. The reaction can be utilized in the derivatization of natural products and drug derivatives, showcasing a high level of tolerance towards functional groups. The authors propose that both the generation of cobalt(III)-alkyl and the SN2-substitution of alkylcobalt(IV) with nucleophiles play a significant role in the formation of enantioselective C–N bonds.
Prompted by these findings, the same group further reported a cobalt-catalyzed intramolecularly enantioselective hydroamination of unactivated amino alkenes (Scheme 32).67 They used a phenanthrene-containing cobalt(salen) complex Co-17 as the catalyst, NFSI as the oxidant and phenylsilane as the hydride source to perform the reaction, delivering a series of chiral pyrrolidines 98, which are key structures in pharmaceuticals and biologically active molecules. This reaction is remarkable in its ability to outcompete two different nucleophilic nitrogen sources within the same system, achieving high chemical selectivity in the conversion of intramolecular sulfonamides. Similar to the previous report, preliminary mechanistic studies demonstrated the involvement of a carbon radical intermediate and suggested that an SN2-type displacement of the CoIV–alkyl intermediate may be operative in the enantioselective transformation.
This catalytic platform has been further expanded very recently to the enantioselective C–O and C–C bond formation by trapping the in situ generated carbon radical intermediates with oxygen and carbon nucleophiles, respectively. It is undeniable that the use of this cobalt catalytic system greatly expands the potential applications of cobalt chemistry and asymmetric radical reactions. It introduces a new method for achieving highly enantioselective coupling of radical intermediates with nucleophiles by trapping the radical intermediate to produce an unusual, electrophilic cationic alkylcobalt(IV) species. This innovative reaction pathway, based on CoIV intermediates and their enantioselective control, represents significant advancements in catalytic synthetic methodology and will find wide application in the future.
In 2008, Zhang and coworkers70 introduced the first use of cobalt(II) complexes with D2-symmetric chiral porphyrins for the asymmetric aziridination of olefins, utilizing diphenylphosphoryl azide (DPPA) as the nitrene source. This catalytic system enabled the formation of N-phosphorylated aziridines in good yields with moderate enantioselectivities. This study marks a noteworthy milestone, as it showcases the first cobalt(II)-catalyzed asymmetric olefin aziridination using DPPA, thus laying a foundation for further advancements in chiral cobalt catalysts aimed at achieving more efficient and enantioselective synthesis of aziridines.
The CoII-based metalloradical system enables the homolytic activation of a wide variety of organic azides with different electronic characteristics, facilitating the stereoselective synthesis of three-membered heterocycles. The researchers expanded the substrate scope to include trichloroethoxysulfonyl azides (TcesN3),71N-fluoroaryl azides (ArFN3),72 and N-carbonyl azides (TrocN3),73 achieving high yields and excellent enantioselectivities (102–104). Mechanistically, this system generates an α-CoIII-aminyl radical intermediate I from organic azides, which then undergoes regioselective addition to olefins, forming a γ-CoIII-alkyl radical intermediate II. This intermediate undergoes an enantioselective 3-exo-tet radical cyclization, resulting in the formation of chiral aziridines 104 (Scheme 33).
In 2017, the Zhang74 group developed an innovative application of cobalt(II)-based MRC for the enantioselective radical bicyclization of allyl azidoformates (Scheme 34). This approach aimed to create a highly efficient and stereoselective catalytic system for constructing aziridine/oxazolidinone-fused bicyclic structures, which are crucial intermediates in the synthesis of biologically relevant compounds like chiral oxazolidinones and vicinal amino alcohols. This method demonstrated broad substrate compatibility, allowing a variety of allyl azidoformates with different substituents (aryl, vinyl, acyl) to undergo radical bicyclization, yielding products in high yields and with excellent stereocontrol. Mechanistic studies, including EPR spectroscopy and the use of both (E)- and (Z)-isomers of allyl azidoformates, indicated a stepwise radical bicyclization pathway that involves α-CoIII-aminyl and γ-CoIII-alkyl radicals. The enantioselective 5-exo-trig cyclization of the α-CoIII-aminyl radical is followed by a diastereoselective 3-exo-tet cyclization of the γ-CoIII-alkyl radical. Notably, the reaction proceeds via diastereoconvergent bicyclization with 1,2-substituted olefins, while diastereospecific bicyclization occurs for 1,1,2-trisubstituted olefins.
![]() | ||
Scheme 34 CoII-based MRC-catalyzed enantioselective radical bicyclization initiated by 5-exo-trig cyclization. |
In 2024, the same group75 developed enantioselective radical N-heterobicyclization using a cobalt(II)-based MRC system. Unlike previous reports, the CoIII-aminyl radical undergoes 6-endo-trig addition to a terminal olefin, forming a γ-CoIII-alkyl radical intermediate, which then cyclizes stereospecifically via 3-exo-tet radical cyclization (Scheme 35). Detailed experimental and computational studies revealed that enantioface-selective radical addition, followed by stereospecific radical substitution, are the key steps in achieving high enantioselectivities.
![]() | ||
Scheme 35 CoII-based MRC-catalyzed enantioselective radical bicyclization initiated by 6-endo-trig cyclization. |
In 2017, the same group77 developed an asymmetric aminofluoroalkylation of arylated alkenes involving radical intermediates. Using sulfonyl chloride as the fluoroalkyl radical source and CuI/CPA (L23) as the SET catalyst, they successfully introduced perfluorobutanyl, trifluoromethyl, difluoroacetyl and difluoromethyl groups into urea-substituted olefins. This method provided a direct route to a variety of chiral β-fluoroalkyl amines 113 (Scheme 37).
The Liu group further expanded the scope of this strategy to other transformations, including di-amination and azido amination,78 aminosilylation,79 and arylamination80 reactions. Mechanistic investigations indicated that the chiral C–N bond formation likely occurs through a carbocation intermediate rather than via a CuIII elimination pathway. This radical-based asymmetric amination strategy provides direct access to a wide range of valuable substituted chiral pyrroles 116–119 (Scheme 38).
This Cu(I)/CPA cooperative system can also be extended to the enantioselective three-component trifluoromethylative amination of alkenes,81 incorporating a convertible hydroxy group as a directing group, using hydrazines and Togni's reagent. The reaction proceeds through a radical-carbocation crossover mechanism. The high levels of stereocontrol in the formation of the asymmetric C–N bond are attributed to hydrogen-bonding and ion-pair interactions between the carbocation intermediate and the CPA. This methodology exhibits broad functional group tolerance, effectively accommodating both electron-rich and electron-deficient substituents on the aromatic ring. Moreover, it successfully incorporates various hydrazines, leading to the formation of enantioenriched diarylmethylamines 122 with an α-tertiary stereocentre (Scheme 39).
Very recently, Liu and coworkers82 developed a Cu/anionic chiral N,N,N-ligand (L26)-catalyzed asymmetric radical carboamination of 1,1-disubstituted alkenes, using readily available alkyl halides and arylamines (Scheme 40). This method provides direct access to value-added chiral α-tertiary N-arylamines 124. The protocol demonstrated broad substrate tolerance, accommodating a wide range of α-aryl-substituted acrylamides, electron-deficient aryl and heteroaryl amines, as well as alkyl halides and sulfonyl chlorides, all with good to excellent enantioselectivities. However, for aromatic amines with electron-neutral or -donating groups, dramatically decreased enantioselectivities were observed. This approach offers a versatile platform for the direct synthesis of chiral α-tertiary N-arylamine building blocks, expanding the toolbox for asymmetric amination reactions.
In 2019, Liu and colleagues83 reported a copper/chiral BOX-catalyzed, site- and enantioselective allylic C–H cyanation of complex alkenes. Notably, when an N–F reagent was used with distal olefins as the starting material, the resulting aminocyanation product was obtained with a 92% yield, although it exhibited only 15% enantioselectivity (see Scheme 41). Despite this low enantioselectivity, this work represents a rare example of enantioselective C–N bond formation via N-centered radical addition, highlighting its potential for further advancements in asymmetric radical chemistry.
In 2019, the Chen and He group84 developed a Cu/chiral BOX-catalyzed enantioselective trifluoromethylative amination of O-homoallyl benzimidates. This methodology displays broad functional group tolerance, efficiently converting substrates bearing both electron-donating and electron-withdrawing groups on the phenyl ring, and exhibiting exceptional compatibility with halogen substituents, including iodine (Scheme 42). Mechanistically, the process involves the formation of alkyl radicals from the addition of CF3 radicals to olefins. These alkyl radicals are enantioselectively captured by a chiral CuII species, leading to the formation of a CuIII intermediate, which then undergoes C–N reductive elimination to generate CF3-containing heterocycles.
The proposed mechanism involves SET between LPO and FeII–N3 to generate an FeIII–N3 species III and alkyl radicals IV. IV abstract the halogen atom from fluoroalkyl iodides to form fluoroalkyl radicals, which subsequently add to the double bond, generating benzylic radicals V. The benzylic radicals then undergo an enantioselective azido group transfer, resulting in chiral organoazides 130. Mechanistic studies, supported by mass spectrometry and computational analysis, indicate that stereoselectivity is achieved through weak non-covalent interactions and steric effects within the chiral environment. This work marks a significant advancement in enantioselective radical azidation and underscores the potential of iron catalysis for further developments in asymmetric radical chemistry (Scheme 43).
Very recently, Feng and coworkers86 developed an iron-catalyzed, asymmetric three-component radical carboazidation of electron-deficient alkenes, including α,β-unsaturated ketones, amides, and phosphine oxides, using C(sp3)–H-containing partners as alkyl radical precursors (Scheme 44). The chiral N,N′-dioxide ligand (L29), designed by their group, efficiently facilitates this radical carboazidation, yielding enantioenriched chiral azide compounds 133 bearing a quaternary stereocenter with good yields and high enantioselectivities. The carbonyl group on the olefins acts as a directing group, enhancing asymmetric azide transfer. This protocol is notable for directly utilizing abundant and cost-effective hydrocarbons as alkylating sources, providing rapid access to valuable chiral alkylazides.
In 2024, Gong and coworkers89 developed a photoinduced Pd-catalyzed three-component enantioselective carboamination of dienes (Scheme 46). A broad range of abundant C(sp3)–H-containing partners, such as toluene-type substrates, ethers, amines, esters, and ketones, served as alkyl radical sources, while free amines, including aliphatic, aromatic, primary, and secondary amines, were used as nitrogen sources. This method enabled the synthesis of chiral allylamines with moderate to excellent enantioselectivities under mild conditions. The efficient construction of chiral allyl amines from cost-effective and readily available chemical raw materials is undoubtedly highly appealing.
![]() | ||
Scheme 46 Photoinduced Pd-catalyzed asymmetric 1,2-carboamination of 1,3-dienes via activation of aliphatic C–H bonds. |
Mechanistically, Pd0 was irradiated by blue LEDs to generate excited Pd0L*, which underwent SET with ArBr, producing an aryl radical and PdIL. The aryl radical then engaged in SET with the C(sp3)–H-containing partners to generate alkyl radicals II. These radicals underwent addition to dienes, forming an allyl radical intermediate III. Radical recombination with PdI resulted in a π-allyl PdII complex, which then underwent regio- and enantioselective nucleophilic substitution by the amine, delivering the final products. The use of bulky aryl bromides was crucial to avoid direct aryl radical addition to the 1,3-diene.
Shortly thereafter, Yang and coworkers90 developed a photoinduced Pd-catalyzed asymmetric 1,2-carboamination of conjugated 1,3-dienes using N-hydroxyphthalimide (138) esters as bifunctional reagents (Scheme 47). This method demonstrated broad substrate compatibility, accommodating NHPI esters derived from primary, cyclic, and noncyclic secondary carboxylic acids, and delivering 1,2-carboamination products 139 with moderate to high enantioselectivities.
All of these achievements showcase the promising development of the catalyzed enantioselective radical amination strategy via asymmetric formation of C–N bonds. However, it remains a dynamic and emerging field, with several issues needing further development: (1) alkyl radicals, such as benzyl, α-carbonyl, and allylic/propargyl radicals, have been effectively utilized for enantioselective radical amination; however, achieving enantio-differentiation between two alkyl groups connected by alkyl radicals remains a significant challenge; (2) although radical amination employing NFSI, N–O reagents, organic azides, dioxazolones, amides, and aniline derivatives as nitrogen sources has been successfully realized, direct enantioselective amination of alkylamines or ammonia remains a considerable obstacle; and (3) most reactions rely on transition metal catalysis for enantioselectivity control. For addressing these challenges, the exploration of new reaction modes as well as innovative chiral catalyst design might be necessary. A new methodology for enantioselective C–N bond formation without employing a metal catalyst is highly desirable. It is expected that this research field will attract widespread attention and make further progress in the near future.
This journal is © the Partner Organisations 2025 |