Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Apr 8.
Published in final edited form as: Chem. 2021 Mar 29;7(4):1120–1134. doi: 10.1016/j.chempr.2021.03.001

Metalloradical activation of carbonyl azides for enantioselective radical aziridination

Xavier Riart-Ferrer 1,3, Peng Sang 2,3, Jingran Tao 2,3, Hao Xu 1, Li-Mei Jin 1, Hongjian Lu 2, Xin Cui 2, Lukasz Wojtas 2, X Peter Zhang 1,4,*
PMCID: PMC8049175  NIHMSID: NIHMS1683065  PMID: 33869888

The bigger picture

Organic azides have been increasingly employed as nitrogen sources for catalytic olefine aziridination due to their ease of preparation and generation of benign N2 as the only byproduct. Among common organic azides, carbonyl azides have not been previously demonstrated as effective nitrogen sources for intermolecular olefin aziridination despite the synthetic utilities of N-carbonyl aziridines. As a new application of metalloradical catalysis, we have developed a catalytic system that can effectively employ the carbonyl azide TrocN3 for highly asymmetric aziridination of alkenes at room temperature. The resulting enantioenriched N-Trocaziridines have been shown as valuable chiral synthons for stereoselective synthesis of other chiral aziridines and various chiral amines. The Co(II)-based metalloradical system, which proceeds with distinctive stepwise radical mechanism, may provide a general method for asymmetric synthesis of chiral aziridines from alkenes with organic azides.

SUMMARY

The carbonyl azide TrocN3 (2,2,2-trichloroethoxycarbonyl azide) is a potent nitrogen radical precursor for radical olefin aziridination via Co(II)-based metalloradical catalysis (MRC). The cobalt(II) complex of D2-symmetric chiral amidoporphyrin 3,5-DitBu-QingPhyrin proves to be an efficient catalyst that can activate TrocN3 at room temperature to aziridinate various styrene derivatives, providing chiral N-carbonyl aziridines in high yields with excellent enantioselectivities. The new Co(II)-based catalytic system can even enable asymmetric aziridination of electron-deficient alkenes, such as methyl and ethyl acrylates. In addition to facile removal of Troc group for generation of unprotected aziridines, the resulting N-Troc-aziridines can be effectively opened by different types of nucleophiles to afford a series of chiral amine derivatives with excellent stereospecificity. Several lines of computational and experimental evidence support the underlying stepwise radical mechanism for Co(II)-catalyzed olefin aziridination. This represents the first example of asymmetric intermolecular olefin aziridination that employs carbonyl azides as the nitrogen source.

Graphical Abstract

graphic file with name nihms-1683065-f0001.jpg

A Co(II)-based system has been developed for enantioselective radical aziridination of alkenes using the carbonyl azide TrocN3 as the nitrogen source. The catalytic system, which operates at room temperature under neutral and nonoxidative conditions, is applicable to both aromatic and electron-deficient olefins and enables the synthesis of N-carbonyl aziridines in high yields with excellent enantioselectivities. The enantioenriched N-carbonyl aziridines have been demonstrated as valuable intermediates for stereoselective organic synthesis. The Co(II)-catalyzed aziridination has been shown to proceed with stepwise radical mechanism.

INTRODUCTION

The development of selective radical reactions for applications in stereoselective organic synthesis has attracted growing research interest.1 While radical reactions have a number of inherent synthetic advantages, such as fast reaction rate and functional group tolerance, their synthetic applications have been hampered by outstanding issues associated with control of reactivity and selectivity, especially in the context of enantioselectivity.2 As stable metalloradicals, cobalt(II) complexes of D2-symmetric chiral amidoporphyrins [Co(D2-Por*)] have emerged as a family of open-shell transition metal catalysts for enantioselective radical transformations through catalytic generation of metal-supported organic radical intermediates via metalloradical catalysis (MRC).35 Specifically, metalloradical catalysts [Co(D2-Por*)] have shown to be particularly effective in activating organic azides to generate the corresponding α-Co(III)-aminyl radicals as key intermediates for catalytic radical aziridination of alkenes,6 producing the smallest three-membered N-heterocycles with effective control of reactivity and enantioselectivity.7,8 While previous reports involved the use of phosphoryl, sulfonyl, and aryl azides,6 we were attracted to the possibility of using carbonyl azides, such as 2,2,2-trichloroethoxycarbonyl azide (TrocN3), for radical olefin aziridination via Co(II)-MRC, especially its asymmetric variant by [Co(D2-Por*)] (Scheme 1). Although [Co(TPP)] (TPP = 5,10,15,20-tetraphenylporphyrin) was previously shown to activate TrocN3, it required elevated temperature and elongated reaction time, as well as the use of high catalyst loading.9 We reasoned that the metalloradical activation of TrocN3 to generate α−2,2,2-trichloroethoxycarbonyl-α-Co(III)-aminyl radical intermediate I could be facilitated by [Co(D2-Por*)] as a result of the putative H-bonding interaction between the carbonyl moiety in the azide and the amide unit of the catalyst (Scheme 1).6d,6f In view of the spin delocalization in the α-carbonylaminyl radical intermediate I, however, it was unclear whether it could function as effective nitrogen-centered radical to proceed radical addition to olefins for generation of γ-Co(III)-alkyl radical intermediate II (Scheme 1). Moreover, in order to form the expected aziridines, the resulting carbon-centered radical in intermediate II must undergo competitive 3-exo-tet radical cyclization over 5-endo-trig radical cyclization,10 which would produce oxazolines after subsequent β-scission. More importantly, could the desired 3-exo-tet radical cyclization proceed in an enantioselective fashion? We envisioned to address these and related issues of reactivity and selectivity through the judicious choice of metalloradical catalysts [Co(D2-Por*)]. If successful, it would offer a new radical protocol for stereoselective synthesis of chiral N-carbonyl aziridines (Scheme 1), which have found synthetic and biological applications (see Figure S1).11

Scheme 1.

Scheme 1.

Proposed mechanism for radical aziridination of alkenes with carbonyl azide TrocN3 via Co(II)-MRC

Organic azides have been increasingly employed as nitrogen sources for catalytic aziridination of alkenes due to their attractive attributes, such as ease of preparation and generation of benign N2 as the only byproduct.3g,11d,12 Among common organic azides, carbonyl azides have not been previously demonstrated as effective nitrogen sources for intermolecular olefin aziridination,7b,13 despite the potential formation of useful aziridines bearing N-carbonyl functionality, which can also be easily deprotected to yield N–H aziridines.11e This underdevelopment is mainly attributed to the well-known challenges associated with the high lability of carbonyl azides toward thermal and photolytic rearrangements.14 Yoon and coworkers recently reported the use of Ir-based catalysts as photosensitizers for activation of TrocN3 with visible light to generate triplet free nitrenes that can be selectively trapped by various alkenes for non-asymmetric aziridination.15 As a novel alternative to carbonyl azides, Lebel and coworkers successfully employed chiral N-tosyloxycarbamates as nitrogen sources for Rh2-catalyzed diastereoselective aziridination of alkenes.16 To the best of our knowledge, there has been no previous report on asymmetric catalytic system for olefin aziridination using TrocN3 as nitrogen source. In comparison with other nitrene precursors, TrocN3 offers several advantages: (1) straightforward synthesis from the commercially available 2,2,2-trichloroethoxy chloroformate in near quantitative yield; (2) generation of environmentally benign dinitrogen as the only byproduct; and (3) ease of deprotection form the resulting aziridine products to afford N–H aziridines. As a new application of Co(II)-MRC, we herein wish to report the development of the first asymmetric catalytic system that can effectively employ TrocN3 for highly asymmetric aziridination of aromatic alkenes. Supported by D2-symmetric chiral amidoporphyrin ligands, the Co(II)-based catalytic process allows for efficient synthesis of chiral N-Troc-aziridines with a high degree of enantiocontrol. In addition to the convenient access of chiral N–H-aziridines through mild deprotection, the utilities of the resulting chiral N-Trocaziridines are further showcased by the synthesis of chiral amines through stereospecific ring-opening with nucleophiles of different nature without erosion of the original enantiopurity. We also describe our mechanistic studies on the proposed stepwise radical mechanism of the Co(II)-catalyzed aziridination.

RESULTS AND DISCUSSION

Reaction optimization

Our efforts started with the use of styrene (2a) as the model substrate for asymmetric aziridination with TrocN3 (1) by metalloradical catalysts [Co(Por)] (Figure 1). While [Co(TPP)] was ineffective (Figure 1; entry 1), the Co(II) complex of D2h-symmetric achiral amidoporphyrin [Co(P1)] (P1 = 3,5-DitBu-IbuPhyrin)6d could catalyze the reaction even at room temperature to form the desired N-Troc-aziridine 3a in a low but significant yield (Figure 1; entry 2), indicating ligand-accelerated catalysis as a result of the putative H-bonding interaction between the carbonyl moiety in the azide and the amide unit of the catalyst. The first-generation chiral metalloradical catalyst [Co(P2)] (P2 = 3,5-DitBu-ChenPhyrin)17 could give rise to significant asymmetric induction for the formation of chiral aziridine 3a while slightly improving the yield (Figure 1; entry 3). Taking advantage of the modularity and tunability of the D2-symmetric chiral amidoporphyrin ligand platform, we synthesized the second-generation metalloradical catalyst [Co(P3)] (P3 = 3,5-DitBu-QingPhyrin) by replacing one of the methyl groups on the chiral amide units of [Co(P2)] with a phenyl group, resulting in a [Co(D2-Por*)] complex bearing chiral amides with two contiguous stereogenic centers.18 Gratifyingly, [Co(P3)] could further enhance both reactivity and enantioselectivity of the catalytic aziridination reaction, affording 3a in 50% yield with 94% ee (Figure 1; entry 4). On the assumption that the presence of adventitious water might negatively affect the yield of the aziridine product, molecular sieves were employed as additives in the catalytic system. Indeed, the yield of aziridine 3a was improved to 81% with preservation of the 94% ee (Figure 1; entry 5). It was further found that addition of anhydrous K2CO3 could lead to quantitative formation of the desired aziridine 3a with the same excellent enantioselectivity (Figure 1; entry 6). While its exact role was difficult to ascertain, we speculated that anhydrous K2CO3 might, in addition to the removal of adventitious water-like molecular sieves, prevent Co(III)-intermediates from functioning as potential Lewis acids to promote aziridine isomerization.19 Switching the ratio of 2a:1 from 3:1 to 1:1.2 resulted in decrease in the yield of 3a without affecting the enantioselectivity (Figure 1; entry 7). Among different solvents tested (see Figure S2), chlorobenzene proved to be the choice of reaction medium, affording aziridine 3a in high yield with excellent enantioselectivity at room temperature (Figure 1; entries 6–10). As expected, no reaction was observed in the absence of a catalyst (Figure 1; entry 11).

Figure 1. Enantioselective aziridination reaction of styrene with carbonyl azide TrocN3 by [Co(Por)].

Figure 1.

aCarried out at room temperature for 24 h with TrocN3 (0.1 mmol) and styrene (0.3 mmol); [TrocN3] = 0.10 M.

b Isolated yields.

cEnantiomeric excess determined by chiral HPLC.

dWith 50 mg of 4 Å molecular sieves.

eWith 0.5 mmol of anhydrous K2CO3.

fTrocN3 (0.12 mmol) and styrene (0.1 mmol) were used.

g Not determined.

Substrate scope

Under the optimized reaction conditions, the [Co(P3)]/TrocN3-based system was found to be effective for enantioselective aziridination of various styrene derivatives (Figure 2). Like styrene, the Co(II)-catalyzed aziridination was suitable for styrene derivatives bearing alkyl substituents at the para-, meta-, and ortho-positions, affording the desired aziridines 3b–3e in moderate to high yields with excellent enantioselectivities (Figure 2; entries 1–4). Fluorinated styrenes having different substitution patterns could be employed as effective substrates, providing the corresponding fluorinated aziridines 3f–3i in high yields with excellent enantioselectivities (Figure 2; entries 5–8). Other halogenated aromatic olefins, such as brominated and chlorinated styrenes, could also be effectively aziridinated with TrocN3, affording the halogenated aziridines 3j–3n in high yields with excellent enantioselectivities (Figure 2; entries 9–13), which may be potentially transformed to other aziridine derivatives by cross-coupling and related reactions. Furthermore, the Co(II)-based system could tolerate functional groups as exemplified by productive formation of the desired aziridine 3o and 3p with high enantioselectivities (Figure 2; entries 14 and 15). To demonstrate the synthetic practicality, a gram-scale synthesis of aziridine 3p was performed using 1 mol % of [Co(P3)] under otherwise the same conditions, affording the desired product with the same high enantioselectivity (92% ee) but in a relatively lower yield (from 71% to 51%). The reduction of the catalyst loading from 2 to 1 mol % is likely responsible for the lower yield. Additionally, styrene derivatives containing electron-withdrawing substituents, such as –CF3 and –NO2 groups at different positions, could serve as suitable substrates for the aziridination system, affording the corresponding three-membered N-heterocycles 3q–3t in enantioenriched forms although in relatively lower yields (Figure 2; entries 16–19). Similarly, the Co(II)-catalyzed aziridination could be applied for styrene derivatives bearing electron-donating groups as demonstrated by the highly asymmetric formation of 3u and 3v albeit in lower yields (Figure 2; entries 20 and 21). In addition, the aziridination system by [Co(P3)] could be applied to extended aromatic olefins as shown by the construction of aziridines 3w–3y in moderate to high yields with excellent enantioselectivities (Figure 2; entries 22–24). The absolute configurations of the newly generated stereogenic centers in both 3p and 3w were established as (R) by X-ray crystallography. Besides mono-substituted styrene derivatives, the catalytic system could be applicable to 1,1-disubstituted styrenes as exemplified by the productive formation of aziridine 3z bearing a tertiary fluoride stereocenter from the reaction of α-fluorostyrene although in lower yield and enantioselectivity (Figure 2; entry 25). Assisted by a catalytic amount of Pd(OAc)2,6e we found that the Co(II)-based metalloradical system could even aziridinate electron-deficient alkenes as shown for the successful reactions of ethyl and methyl acrylates to form the corresponding aziridines 3aa and 3ab, respectively, in moderate yields but with significant levels of enantiocontrol (Figure 2; entries 26–27). It is worth mentioning that electron-deficient alkenes are known to be challenging substrates for aziridination by existing catalytic systems involving electrophilic metallonitrenes as the key intermediate. When heteroaromatic olefins were used, however, it generated an unidentified mixture of products in low yields without observation of the corresponding aziridines. Catalyst [Co(P3)] were found to be ineffective for aziridination reactions of aliphatic and internal olefins with TrocN3, as well as dienes. Evidently, a more effective catalyst needs to be developed for asymmetric aziridination with a broader scope of alkenes.

Figure 2. Co(II)-catalyzed enantioselective aziridination of styrene derivatives with carbonyl azide TrocN3.

Figure 2.

aCarried out with 1 (0.1 mmol), 2 (0.3 mmol), and K2CO3 (0.5 mmol) by [Co(P3)] (2 mol %) in chlorobenzene at room temperature for 24 h; [TrocN3] = 0.10 M; Isolated yields; Enantiomeric excess determined by chiral HPLC.

b Used [Co(P3)] (5 mol %), 2 (0.5 mmol), and Cs2CO3 (0.5 mmol).

cAt 0°C for 24 h.

dAt 40°C for 24 h.

e Absolute configuration determined by anomalous dispersion effects in X-ray diffraction measurements on the crystal.

f51% Yield (815 mg); 92% ee when carried out on a gram scale with 1 (4.6 mmol), 2 (13.7 mmol), and K2CO3 (22.9 mmol) with [Co(P3)] (1 mol %).

gAt −20°C for 48 h.

hUsed [Co(P3)] (5 mol %) with addition of Pd(OAc)2 (10 mol %) in the presence of 4 Å molecular sieves (50 mg) instead of K2CO3 at 40°C for 48 h.

Mechanistic studies

Combined computational and experimental studies were performed for the proposed stepwise radical mechanism of the Co(II)-catalyzed aziridination (Scheme 1). First, density functional theory (DFT) calculations were carried out to study the catalytic pathway for aziridination reaction of styrene with TrocN3 with the use of the actual catalyst [Co(P3)] (Scheme 2A).20 The computational study indicates the existence of intermediate B, which is formed upon coordination of the internal nitrogen atom in TrocN3 to the cobalt center of the catalyst. The formation of intermediate B is slightly exergonic by −1.6 kcal/mol due to additional H-bonding stabilization (see Schemes S6 and S7). Upon further activation, the coordinated azide undergoes dinitrogen elimination to generate α-Co(III)-aminyl radical C. The metalloradical activation step, which is exergonic by −12.2 kcal/mol, has a relative low activation energy (TS1: ΔG = 14.0 kcal/mol). As shown in the middle of the catalytic cycle (see Schemes S6 and S7), the optimized TS1 structure reveals strong double H-bond interactions (N–H—N: 2.34 Å; N–H—O: 2.22 Å) between [Co(P3)] and TrocN3, as well as the strengthening of Co–N bonding interaction (as indicated by the decrease of bond distance from 2.10 to 1.90 Å). According to the DFT calculation, the subsequent radical addition of radical intermediate C to styrene is associated with a relatively high but accessible activation barrier (TS2: ΔG = 21.0 kcal/mol), leading to the formation of γ-Co(III)-alkyl radical intermediate D (Scheme 2A). The final step of 3-exo-tet cyclization via radical substitution, which is highly exergonic by –27.2 kcal/mol, is found to be almost barrierless, resulting in the formation of the three-membered aziridine 3a and the regeneration of catalyst [Co(P3)]. For all computational details including cartesian coordinates of intermediates and transition states, see Data S1.

Scheme 2. Mechanistic studies for the [Co(P3)] catalyzed aziridination of styrene with TrocN3.

Scheme 2.

aAll relative Gibbs free energies (ΔG°) for intermediates and transition states, as well as activation energies (ΔG), are reported in kcal/mol.

b Carried out with 1 (0.1 mmol), 2aD (0.3 mmol), and K2CO3 (0.5 mmol) by [Co(Por)] (2 mol %) in chlorobenzene at 40 °C for 24 h; [TrocN3] = 0.10 M; (Z)-3aD:(E)-3aD ratio determined by 1H-NMR and 2H-NMR analysis of crude reaction mixture; see Supplemental information for details.

Second, in an effort to detect the α-Co(III)-aminyl radical intermediate I experimentally, the isotropic X-band electron paramagnetic resonance (EPR) spectrum was recorded at room temperature for the reaction mixture of [Co(P3)] with TrocN3 (1) in benzene without alkene substrate (Scheme 2B). The spectrum displays salient signals akin to those characteristics of α-Co(III)-aminyl radicals (Scheme 2B).8,21 The observed isotropic g value of 2.00 is consistent with the formation of organic radical I[Co(P3)] upon spin translocation from the Co(II) to the N-atom during metalloradical activation of the azide. In consistence with the spin delocalization in the α-carbonylaminyl radical intermediate I[Co(P3)], the observed broad signals could be fittingly simulated by invoking its three resonance forms on the basis of hyperfine couplings by both 14N (I = 1) and 59Co (I = 7/2): 17% of N-centered radical NI[Co(P3)] (g: 2.00236; A(Co): 8.9 MHz; A(N): 50.9 MHz), 78% of C-centered radical CI[Co(P3)] (g: 2.00931; A(Co): 7.5 MHz; A(N): 46.9 MHz), and 5% of O-centered radical OI[Co(P3)] (g: 2.00206; A(Co): 0 MHz; A(N): 35.7 MHz) (see the Supplemental information for details). Furthermore, the α-Co(III)-aminyl radical I[Co(P3)] from the reaction mixture of [Co(P3)] with azide 1 could be detected by high-resolution mass spectrometry (HRMS) with electrospray ionization (ESI) ionization (see the Supplemental information for details). The obtained spectrum (Scheme 2B) evidently exhibited a signal corresponding to [I[Co(P3)]]+ (m/z = 1,776.6666), which resulted from the neutral α-Co(III)-aminyl radical I[Co(P3)] by the loss of one electron. Both the exact mass and the isotope distribution pattern measured experimentally matched well with those calculated from the formula of [(P3)Co(NCO2CH2CCl3)]+. The successful detection of α-Co(III)-aminyl radical intermediate I[Co(P3)] by EPR and HRMS provided experimental evidence to support the first step of metalloradical activation in the proposed mechanism of radical aziridination (Scheme 1).

Finally, to probe the subsequent steps of radical addition and radical cyclization in the proposed mechanism (Scheme 1), both isotopomers of β-deuterostyrene (E)-2aD and (Z)-2aD were applied as substrates for Co(II)-catalyzed aziridination with TrocN3 (1) (Scheme 2C). While a concerted mechanism associated with metallonitrene intermediates is usually stereospecific, a stepwise mechanism involving α-Co(III)-aminyl radical intermediates may lead to the formation of aziridines as a mixture of both diastereoisomers (E)-3aD and (Z)-3aD from either (E)-2aD or (Z)-2aD as a result of the β-C–C bond rotation in the resulting γ-Co(III)-alkyl radical intermediate II before cyclization (Scheme 2C; Schemes S1 and S2). As expected, the aziridination of (E)-2aD with azide 1 by the achiral catalyst [Co(P1)] produced an isotopomeric mixture of (Z)-3aD and (E)-3aD in a ratio of 20:80. Under the identical conditions, aziridination of (Z)-2aD yielded the isotopomeric mixture in a ratio of 89:11. The formation of isotopomeric mixture of aziridine 3aD from isomerically pure 2aD is clearly attributed to the rotation around the β-C–C bond in the γ-Co(III)-alkyl radical intermediate II. Interestingly, it was found that the degree of the bond rotation could be influenced by the environment of the supporting ligand. For example, when the two reactions were carried out under the same condition but using the chiral catalyst [Co(P2)], the isotopomeric ratio of (Z)-3aD and (E)-3aD changed from 20:80 to 12:88 for the reaction of (E)-2aD and from 89:11 to 94:06 for the reaction of (Z)-2aD, indicating a lower degree of rotation in a more hindered ligand environment (Scheme 2C). Accordingly, when the optimized catalyst [Co(P3)], which has an even more confined ligand environment, was employed for the aziridination reactions of (E)-2aD and (Z)-2aD, minimal isotopomeric distributions were observed as a result of faster radical cyclization than the β-C–C bond rotation, leading to a stereospecific process (Scheme 2C). This high stereospecificity for β-deuterostyrenes, together with the high enantioselectivity observed for styrene (Figure 1, entry 6), implies that [Co(P3)]-catalyzed olefin aziridination with TrocN3 involves radical addition as the enantiodetermining step, followed by a stereoretentive radical cyclization. As anticipated for a radical process, it was found that the catalytic aziridination process could be significantly inhibited when a large excess of TEMPO (2,2,6,6-Tetramethylpiperidine 1-oxyl) was added. Together with the detection of the α-Co(III)-aminyl radical intermediate I by EPR and HRMS, these results fully corroborate the proposed stepwise radical mechanism for the Co(II)-based metalloradical aziridination (Scheme 1).

Synthetic applications

For access to aziridine derivatives with various N-substituents for different applications, it would be desirable that the aziridine products from a catalytic aziridination process could be effectively converted to the corresponding N–H aziridines through simple deprotection without opening the three-membered ring structures.11e,22 In contrast to aziridines with N-protecting groups that require harsh conditions for deprotection,23 N-carboalkoxy aziridines have been known to proceed facile N-deprotection under mild conditions.24 To showcase the synthetic utilities of the resulting chiral N-Troc-aziridines from the Co(II)-catalyzed process, it was demonstrated that enantioenriched N-Troc-aziridine (R)-3p could be readily converted to the corresponding N–H aziridine 4p in 87% yield without erosion of its optical purity when treated with lithium hydroxide at room temperature (Scheme 3A). In addition to N-deprotection, chiral N-Troc-aziridines were shown to undergo facile ring-opening reactions at room temperature by a wide range of nucleophiles in the presence of Lewis acids with preservation of the high enantiomeric purity (Scheme 3A). For instance, methanol could effectively open the three-membered ring in (R)-3p in the presence of boron trifluoride diethyl etherate, generating chiral β-amino ether 5p in 96% yield without any racemization. Notably, even water could function as an effective nucleophile for ring-opening of (R)-3p under similar conditions, resulting in highly stereospecific formation of chiral b-amino alcohol 6p in 94% yield. The absolute configuration of 6p was established by X-ray crystallography as (S), indicating an SN2-type mechanism of the ring-opening reaction. Sulfur-based nucleophiles could also be applied for the ring-opening process as exemplified by the reaction of (R)-3p with thiophenol, affording the chiral β-amino thioether 7p in 58% yield with some loss of the original enantiopurity. The three-membered ring in chiral N-Troc-aziridines could be readily opened by nitrogen-based nucleophiles, including both aromatic and aliphatic amines at room temperature to provide valuable vicinal diamines.25 For example, both aniline and diethylamine could effectively react with (R)-3p to afford chiral vicinal diamines 8p and 9p, respectively, in high yields with no erosion of the original enantiopurity. Interestingly, when primary aliphatic amines were employed as the nucleophiles, the reactions of N-Troc-aziridines were found to proceed further to generate aziridine-based chiral ureas, a formal process that transforms carbamates to ureas via amide bond formation (Scheme 3B).26 As two examples, benzylamine and allylamine reacted readily with (R)-3p under the similar conditions to form aziridine-based chiral ureas 10p and 11p, respectively, in excellent yields without apparent racemization. Presumably due to the combined high nucleophilicity of the resulting secondary amines and good leaving ability of trichloroethyl group, the initial ring-opening products III of (R)-3p by the primary amines proceeded further to form intramolecular amide bonds in the presence of Lewis acids to generate imidazolidinones IV, which then underwent ring-contraction under Lewis acid catalysis27 to yield the final products 10p and 11p. The N-deprotection and ring-opening reactions could proceed equally well with other chiral N-Troc-aziridines as exemplified by the stereospecific formations of N–H aziridine 4r and β-amino ether 5r from enantioenriched aziridine 3r (Scheme 3C).

Scheme 3. Transformations of chiral N-Troc-aziridines.

Scheme 3.

aConditions as reported in the literature.

bMeOH (0.10 M), BF3•OEt2 (10 mol %).

cH2O (0.10 M), BF3•OEt2 (10 mol %).

dPhSH (0.10 M), BF3•OEt2 (10 mol %).

ePhNH2 (0.10 M), BF3•OEt2 (10 mol %).

fCH2Cl2 (0.10 M), BF3•OEt2 (20 mol %), amines (20 equiv).

Conclusions

In summary, we have developed the first catalytic system that can employ the carbonyl azide TrocN3 as the nitrogen source for asymmetric aziridination via Co(II)-based MRC. With the support of 3,5-DitBu-QingPhyrin as the chiral ligand, the Co(II)-based aziridination system with TrocN3, which proceeds with the underlying stepwise radical mechanism as evidenced by the combined computational and experimental studies, provides a fundamentally new methodology for stereoselective synthesis of chiral N-Troc-aziridines from styrene derivatives with varied steric and electronic properties in high yields with high enantioselectivities. Assisted by a catalytic amount of Pd(OAc)2, this new [Co(3,5-DitBu-QingPhyrin)]/TrocN3-based system can further catalyze asymmetric aziridination of electron-deficient alkenes, such as methyl and ethyl acrylates, which are challenging substrates for the catalytic systems involving electrophilic metallonitrene intermediates. Among several salient features, the Co(II)-catalyzed radical aziridination can operate efficiently at room temperature and has good functional group tolerance. The resulting enantioenriched N-Troc-aziridines have been showcased as valuable chiral synthons for stereoselective synthesis of other chiral aziridines and various chiral amine derivatives. Through fine-tuning the environments of D2-symmetric chiral amidoporphyrins as the supporting ligand, we hope to develop a more general Co(II)-based catalytic system for enantioselective radical aziridination of different alkenes with TrocN3.

EXPERIMENTAL PROCEDURES

Resource availability

Lead contact

Further information and requests for resources should be directed to and will be fulfilled by the lead contact, X. Peter Zhang (peter.zhang@bc.edu).

Materials availability

Unique and stable reagents generated in this study will be made available on request, but we might require a payment and/or a completed materials transfer agreement if there is potential for commercial application.

Data and code availability

The crystal structure data of compound (R)-3p, (R)-3w, and (S)-6p have been deposited in the Cambridge structural database under reference numbers CCDC: 2025803, 2025804, and 2025805, respectively.

Full experimental procedures are provided in the Supplemental information.

Supplementary Material

1
2

Highlights.

Catalytic system for enantioselective aziridination of alkenes with TrocN3

Highly enantioselective synthesis of chiral N-Troc-aziridines at room temperature

Systematic demonstration of chiral N-Troc-aziridines for stereoselective synthesis

Detailed mechanistic studies on elucidation of underlying stepwise radical pathway

ACKNOWLEDGMENTS

We are grateful for financial support by NIH (R01-GM102554) and, in part, by NSF (CHE-1900375).

Footnotes

DECLARATION OF INTERESTS

The authors declare no competing interests.

SUPPLEMENTAL INFORMATION

Supplemental information can be found online at https://doi.org/10.1016/j.chempr.2021.03.001.

REFERENCES

  • 1.For select books, see:Chatgilialoglu C, and Studer A (2012). Encyclopedia of Radicals in Chemistry, Biology and Materials (John Wiley & Sons; ).Curran DP, Porter NA, and Giese B (2008). Stereochemistry of Radical Reactions: Concepts, Guidelines, and Synthetic Applications (John Wiley & Sons; ).Zard SZ (2003). Radical Reactions in Organic Synthesis (Oxford University Press; ), For recent reviews, see: Zard SZ (2019). Radical alliances: solutions and opportunities for organic synthesis. Helv. Chim. Acta 102, e1900134.Huang HM, Garduño-Castro MH, Morrill C, and Procter DJ (2019). Catalytic cascade reactions by radical relay. Chem. Soc. Rev 48, 4626,–4638.Studer A, and Curran DP (2016). Catalysis of radical reactions: a radical chemistry perspective. Angew. Chem. Int. Ed. Engl 55, 58,–102.Brimioulle R, Lenhart D, Maturi MM, and Bach T (2015). Enantioselective catalysis of photochemical reactions. Angew. Chem. Int. Ed. Engl 54, 3872,–3890.Prier CK, Rankic DA, and MacMillan DWC (2013). Visible light photoredox catalysis with transition metal complexes: applications in organic synthesis. Chem. Rev 113, 5322,–5363.Narayanam JMR, and Stephenson CRJ (2011). Visible light photoredox catalysis: applications in organic synthesis. Chem. Soc. Rev 40, 102,–113.Quiclet-Sire BZ, and Zard SZ (2011). Fun with radicals: some new perspectives for organic synthesis. Pure Appl. Chem 83, 519,–551.Zard SZ (2008). Recent progress in the generation and use of nitrogen-centred radicals. Chem. Soc. Rev 37, 1603–1618. [Google Scholar]
  • 2.For select examples on approaches to control radical reactivity and selectivity, see:Du J, Skubi KL, Schultz DM, and Yoon TP (2014). A dual-catalysis approach to enantioselective [2 + 2] photocycloadditions using visible light. Science 344, 392–396.Huo H, Shen X, Wang C, Zhang L, Röse P, Chen LA, Harms K, Marsch M, Hilt G, and Meggers E (2014). Asymmetric photoredox transition-metal catalysis activated by visible light. Nature 515, 100,–103.Lin JS, Dong XY, Li TT, Jiang NC, Tan B, and Liu XY (2016). A dual-catalytic strategy to direct asymmetric radical Aminotrifluoromethylation of alkenes. J. Am. Chem. Soc 138, 9357,–9360.Kainz QM, Matier CD, Bartoszewicz A, Zultanski SL, Peters JC, and Fu GC (2016). Asymmetric copper-catalyzed C–N cross-couplings induced by visible light. Science 351, 681,–684.Zhang W, Wang F, McCann SD, Wang D, Chen P, Stahl SS, and Liu G (2016). Enantioselective cyanation of benzylic C–H Bonds Via copper-catalyzed radical relay. Science 353, 1014,–1018.Funken N, Mühlhaus F, and Gansäuer A (2016). General, highly selective synthesis of 1,3- and 1,4-Difunctionalized building blocks by regiodivergent epoxide opening. Angew. Chem. Int. Ed. Engl 55, 12030,–12034.Brill ZG, Grover HK, and Maimone TJ (2016). Enantioselective synthesis of an ophiobolin sesterterpene via a programmed radical cascade. Science 352, 1078,–1082.Kern N, Plesniak MP, McDouall JJW, and Procter DJ (2017). Enantioselective cyclizations and cyclization cascades of samarium ketyl radicals. Nat. Chem 9, 1198,–1204.Morrill C, Jensen C, Just-Baringo X, Grogan G, Turner NJ, and Procter DJ (2018). Biocatalytic conversion of cyclic ketones bearing a-quaternary stereocenters into lactones in an enantioselective radical approach to mediumsized carbocycles. Angew. Chem. Int. Ed. Engl 57, 3692,–3696.Huang H-M, McDouall JJW, and Procter DJ (2019). SmI2-catalysed cyclization cascades by radical relay. Nat. Catal 2, 211,–218.Ye L, Tian Y, Meng X, Gu Q-S, and Liu X-Y (2020). Enantioselective copper(I)/chiral phosphoric acid catalysed intramolecular amination of allylic and benzylic C–H bonds. Angew. Chem. Int. Ed. Engl 59, 1129,–1133.Li X-T, Lv L, Wang T, Gu Q-S, Xu G-X, Li Z-L, Ye L, Zhang X, Cheng G-J, and Liu X-Y (2020). Diastereo- and enantioselective catalytic radical oxysulfonylation of alkenes in β,γ-unsaturated ketoximes. Chem 6, 1692,–1706.Roos CB, Demaerel J, Graff DE, and Knowles RR (2020). Enantioselective hydroamination of alkenes with sulfonamides enabled by protoncoupled electron transfer. J. Am. Chem. Soc. 142, 5974,–5979.Perego LA, Bonilla P, and Melchiorre P (2020). Photo-organocatalytic enantioselective radical cascade enabled by single-electron transfer activation of allenes. Adv. Synth. Catal 362, 302,–307.Cheng Y-F, Liu J-R, Gu Q-S, Yu Z-L, Wang J, Li Z-L, Bian J-Q, Wen H-T, Wang X-J, Hong X, et al. (2020). Catalytic enantioselective desymmetrizing functionalization of alkyl radicals via Cu(I)/CPA cooperative catalysis. Nat. Catal 3, 401–410. [Google Scholar]
  • 3.For select reviews and highlights on Co(II)-based MRC, see Huang et al.,1e and: Demarteau J, Debuigne A, and Detrembleur C (2019). Organocobalt complexes as sources of carbon-centered radicals for organic and polymer chemistries. Chem. Rev 119, 6906–6955.Singh R, and Mukherjee A (2019). Metalloporphyrin catalyzed C–H amination. ACS Catal 9, 3604,– 3617.Pellissier H, and Clavier H (2014). Enantioselective cobalt-catalyzed transformations. Chem. Rev 114, 2775,–2823.Lu H, and Zhang XP (2011). Catalytic C–H functionalization by metalloporphyrins: recent developments and future directions. Chem. Soc. Rev 40, 1899,–1909.Che CM, Lo VK-Y, Zhou CY, and Huang JS (2011). Selective functionalisation of saturated C–H bonds with metalloporphyrin catalysts. Chem. Soc. Rev 40, 1950,–1975.Driver TG (2010). Recent advances in transition metal-catalyzed N-atom transfer reactions of azides. Org. Biomol. Chem 8, 3831,–3846.Fantauzzi S, Caselli A, and Gallo E (2009). Nitrene transfer reactions mediated by metallo-porphyrin complexes. Dalton Trans 28, 5434,–5443.Doyle MP (2009). Exceptional selectivity in cyclopropanation reactions catalyzed by chiral cobalt(II)–porphyrin catalysts. Angew. Chem. Int. Ed. Engl 48, 850–852. [Google Scholar]
  • 4.For select examples of Ti(III)-based radicalprocesses, see:Nugent WA, and Rajanbabu TV (1988). Transition-metal-centered radicals in organic synthesis. Titanium(III)-induced cyclization of epoxy olefins. J. Am. Chem. Soc 110, 8561–8562.Rajanbabu TV, and Nugent WA (1994). Selective generation of free radicals from epoxides using a transition-metal radical. A powerful new tool for organic synthesis. J. Am. Chem. Soc 116, 986,–997.Gansäuer A, Rinker B, Pierobon M, Grimme S, Gerenkamp M, and Mück-Lichtenfeld C (2003). A radical tandem reaction with homolytic cleavage of a ti–O bond. Angew. Chem. Int. Ed. Engl 42, 3687,–3690.Gansäuer A, Fan CA, Keller F, and Keil J (2007). Titanocene-catalyzed regiodivergent epoxide openings. J. Am. Chem. Soc 129, 3484,–3485.Gansäuer A, Fleckhaus A, Lafont MA, Okkel A, Kotsis K, Anoop A, and Neese F (2009). Catalysis via homolytic substitutions with C–O and Ti–O bonds: oxidative additions and reductive eliminations in single electron steps. J. Am. Chem. Soc 131, 16989,–16999.Gansäuer A, Hildebrandt S, Michelmann A, Dahmen T, von Laufenberg D, Kube C, Fianu GD, and Flowers RA (2015). Cationic titanocene(III) complexes for catalysis in single-electron steps. Angew. Chem. Int. Ed. Engl 54, 7003,–7006.Gansäuer A, Hildebrandt S, Vogelsang E, and Flowers RA (2016). Tuning the redox properties of the titanocene(III)/(IV)-couple for atom-economical catalysis in single electron steps. Dalton Trans 45, 448,–452.Hao W, Wu X, Sun JZ, Siu JC, MacMillan SN, and Lin S (2017). Radical redox-relay catalysis: formal [3+2] cycloaddition of N-Acylaziridines and alkenes. J. Am. Chem. Soc 139, 12141,– 12144.Klare S, Gordon JP, Gansäuer A, Rajanbabu TV, and Nugent WA (2019). The reaction of b,g-epoxy alcohols with titanium(III) reagents. A proposed role for intramolecular hydrogen bonding. Tetrahedron 75, 130662,.McCallum T, Wu X, and Lin S (2019). Recent advances in titanium radical redox catalysis. J. Org. Chem 84, 14369,–14380.Yao C, Dahmen T, Gansäuer A, and Norton J (2019). Anti-Markovnikov alcohols via epoxide hydrogenation through cooperative catalysis. Science 364, 764,–767.Ye KY, McCallum T, and Lin S (2019). Bimetallic radical redox-relay catalysis for the isomerization of epoxides to allylic alcohols. J. Am. Chem. Soc 141, 9548–9554. [Google Scholar]
  • 5.For select examples of catalytic radical processes involving metalloradical intermediates, see: Smith DM, Pulling ME, and Norton JR (2007). Tin-free and catalytic radical cyclizations. J. Am. Chem. Soc 129, 770–771.Estes DP, Norton JR, Jockusch S, and Sattler W (2012). Mechanisms by which alkynes react with CpCr(CO)3H. Application to radical cyclization. J. Am. Chem. Soc 134, 15512,–15518.Li G, Han A, Pulling ME, Estes DP, and Norton JR (2012). Evidence for formation of a Co–H bond from (H2O)2Co(dmgBF2)2 under H2: application to radical cyclizations. J. Am. Chem. Soc 134, 14662,–14665.Kuo JL, Hartung J, Han A, and Norton JR (2015). Direct generation of oxygen-stabilized radicals by H• transfer from transition metal hydrides. J. Am. Chem. Soc 137, 1036–1039. [Google Scholar]
  • 6.For intermolecular alkene aziridination, see:Caselli A, Gallo E, Ragaini F, Ricatto F, Abbiati G, and Cenini S (2006). Chiral porphyrin complexes of cobalt(II) and ruthenium(II) in catalytic cyclopropanation and amination reactions. Inorg. Chim. Acta 359, 2924–2932.Gao GY, Jones JE, Vyas R, Harden JD, and Zhang XP (2006). Cobaltcatalyzed aziridination with diphenylphosphoryl azide (DPPA): direct synthesis of N-phosphorus-substituted aziridines from alkenes. J. Org. Chem 71, 6655,–6658.Jones JE, Ruppel JV, Gao GY, Moore TM, and Zhang XP (2008). Cobaltcatalyzed asymmetric olefin aziridination with diphenylphosphoryl azide. J. Org. Chem 73, 7260,–7265.Ruppel JV, Jones JE, Huff CA, Kamble RM, Chen Y, and Zhang XP (2008). A highly effective cobalt catalyst for olefin aziridination with azides: hydrogen bonding guided catalyst design. Org. Lett 10, 1995,– 1998.Subbarayan V, Ruppel JV, Zhu S, Perman JA, and Zhang XP (2009). Highly asymmetric cobalt-catalyzed aziridination of alkenes with Trichloroethoxysulfonyl azide (TcesN3). Chem. Commun. (Camb) 2009, 4266,– 4268.Jin LM, Xu X, Lu H, Cui X, Wojtas L, and Zhang XP (2013). Effective synthesis of chiral N-Fluoroaryl aziridines through enantioselective aziridination of alkenes with fluoroaryl azides. Angew. Chem. Int. Ed. Engl 52, 5309,–5313.Tao J, Jin LM, and Zhang XP (2014). Synthesis of chiral N-phosphoryl aziridines through enantioselective aziridination of alkenes with phosphoryl azide via Co(II)-based metalloradical catalysis. Beilstein J. Org. Chem 10, 1282,–1289.Subbarayan V, Jin LM, Xin C, and Zhang XP (2015). Room temperature activation of aryloxysulfonyl azides by [Co(II)(TPP)] for selective radical aziridination of alkenes via metalloradical catalysis. Tetrahedron Lett 56, 3431,–3434.Fantauzzi S, Gallo E, Caselli A, Piangiolino C, Ragaini F, and Cenini S (2007). The (porphyrin) ruthenium-catalyzed aziridination of olefins using aryl azides as nitrogen sources. Eur. J. Org. Chem 6053,–6059.Caselli A, Gallo E, Fantauzzi S, Morlacchi S, Ragaini F, and Cenini S (2008). Allylic amination and aziridination of olefins by aryl azides catalyzed by CoII (tpp): a synthetic and mechanistic study. Eur. J. Inorg. Chem 3009,–3019.Zardi P, Pozzoli A, Ferretti F, Manca G, Mealli C, and Gallo E (2015). A mechanistic investigation of the ruthenium porphyrin catalysed aziridination of olefins by aryl azides. Dalton Trans 44, 10479–10489. [Google Scholar]
  • 7.For intramolecular alkene aziridination, see: Jiang H, Lang K, Lu H, Wojtas L, and Zhang XP (2016). Intramolecular radical aziridination of allylic sulfamoyl azides by cobalt(II)-based metalloradical catalysis: effective construction of strained heterobicyclic structures. Angew. Chem. Int. Ed. Engl 55, 11604–11608.Jiang H, Lang K, Lu H, Wojtas L, and Zhang XP (2017). Asymmetric radical Bicyclization of allyl azidoformates via cobalt(II)-based metalloradical catalysis. J. Am. Chem. Soc 139, 9164,–9167.Alderson JM, Corbin JR, and Schomaker JM (2017). Tunable, chemo- and site-selective nitrene transfer reactions through the rational design of silver(I) catalysts. Acc. Chem. Res 50, 2147,–2158.Ju M, Weatherly CD, Guzei IA, and Schomaker JM (2017). Chemo- and enantioselective intramolecular silver-catalyzed aziridinations. Angew. Chem. Int. Ed. Engl 56, 9944,–9948.Corbin JR, Ketelboeter DR, Fernandez I, and Schomaker JM (2020). Biomimetic 2-imino-Nazarov cyclizations via Eneallene aziridination. J. Am. Chem. Soc 142, 5568–5573. [Google Scholar]
  • 8.For radical mechanism of [Co(D2-Por*)]-catalyzed aziridination and related amination, see:Jiang et al.,7b and references therein.Hu Y, Lang K, Li C, Gill JB, Kim I, Lu H, Fields KB, Marshall M, Cheng Q, Cui X, et al. (2019). Enantioselective radical construction of 5-membered cyclic sulfonamides by metalloradical C–H amination. J. Am. Chem. Soc 141, 18160–18169.Goswami M, Lyaskovskyy V, Domingos SR, Buma WJ, Woutersen S, Troeppner O, Ivanovic-Burmazovic I, Lu H, Cui X, Zhang XP, et al. (2015). Characterization of porphyrin-Co(III)-’nitrene radical’ species relevant in catalytic nitrene transfer reactions. J. Am. Chem. Soc 137, 5468–5479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Lu H, Subbarayan V, Tao J, and Zhang XP (2010). Cobalt(II)-catalyzed intermolecular benzylic CH amination with 2,2,2-trichloroethoxycarbonyl azide (TrocN3). Organometallics 29, 389–393. [Google Scholar]
  • 10.Cui X, Xu X, Wojtas L, Kim MM, and Zhang XP (2012). Regioselective synthesis of multisubstituted furans via metalloradical cyclization of alkynes with a-diazocarbonyls: construction of functionalized a-oligofurans. J. Am. Chem. Soc 134, 19981–19984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Watson IDG, Yu L, and Yudin AK (2006). Advances in nitrogen transfer reactions involving aziridines. Acc. Chem. Res. 39, 194–206.Pellissier H (2010). Recent developments in asymmetric aziridination. Tetrahedron 66, 1509,–1555.Jiang H, and Zhang XP (2012). Oxidation: C–N bond formation by oxidation (aziridines). In Comprehensive Chirality, Carreira EM and Yamamoto H, eds. (Elsevier; ), pp. 168,–182.Degennaro L, Trinchera P, and Luisi R (2014). Recent advances in the stereoselective synthesis of aziridines. Chem. Rev. 114, 7881,–7929.Jat JL, Paudyal MP, Gao H, Xu QL, Yousufuddin M, Devarajan D, Ess DH, Kürti L, and Falck JR (2014). Direct stereospecific synthesis of unprotected N–H and N–me aziridines from olefins. Science 343, 61–65. [Google Scholar]
  • 12.Bräse S, Gil C, Knepper K, and Zimmermann V (2005). Organic azides: an exploding diversity of a unique class of compounds. Angew. Chem. Int. Ed. Engl 44, 5188–5240.Katsuki T (2005). Azide compounds: nitrogen sources for atomefficient and ecologically benign nitrogenatom-transfer reactions. Chem. Lett. 34, 1304–1309. [DOI] [PubMed] [Google Scholar]
  • 13.For select examples on non-asymmetric intramolecular alkene aziridination with carbonyl azides, see: Rhouati S, and Bernou A (1989). Cyclisation of azidoformates, formation of aziridines. J. Chem. Soc. Chem. Commun. 730–732. Bergmeier SC, and Stanchina DM (1997). Synthesis of vicinal amino alcohols via a tandem acylnitrene aziridinationaziridine ring opening. J. Org. Chem 62, 4449,–4456. Bergmeier SC, and Stanchina DM (1999). Acylnitrene route to vicinal amino alcohols. Application to the synthesis of (−)-bestatin and analogues. J. Org. Chem 64, 2852,–2859.Kan C, Long CM, Paul M, Ring CM, Tully SE, and Rojas CM (2001). Photo amidoglycosylation of an Allal azidoformate. Synthesis of β−2-amido allopyranosides. Org. Lett 3, 381,–384. Yoshimitsu T, Ino T, and Tanaka T (2008). Total synthesis of (−)-agelastatin. A. Org. Lett 10, 5457,–5460. Chang YJ, Hsuan YC, Lai AC-Y, Han YC, and Hou DR (2016). Synthesis of α-C-galactosylceramide via diastereoselective aziridination: the new immunostimulant 4′-epi-C-glycoside of KRN7000. Org. Lett 18, 808,–811. Zhang Y, Dong X, Wu Y, Li G, and Lu H (2018). Visible-light-induced intramolecular C(sp2)-H amination and aziridination of azidoformates via a triplet nitrene pathway. Org. Lett 20, 4838–4842. [Google Scholar]
  • 14.Curtius T (1890). Ueber Stickstoffwasserstoffsäure (Azoimid) N3H. Ber. Dtsch. Chem. Ges 23, 3023–3033. Smith PA (1946). The Curtius reaction. Org. React 337–449. [Google Scholar]
  • 15.Scholz SO, Farney EP, Kim S, Bates DM, and Yoon TP (2016). Spin-selective generation of triplet nitrenes: olefin aziridination through visible-light photosensitization of azidoformates. Angew. Chem. Int. Ed. Engl 55, 2239–2242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Lebel H, Spitz C, Leogane O, Trudel C, and Parmentier M (2011). Stereoselective rhodium-catalyzed amination of alkenes. Org. Lett 13, 5460–5463.Lebel H, Parmentier M, Leogane O, Ross K, and Spitz C (2012). Copper bis(oxazolines) as catalysts for stereoselective aziridination of styrenes with N-Tosyloxycarbamates. Tetrahedron 68, 3396,–3409.Azek E, Spitz C, Ernzerhof M, and Lebel H (2020). A mechanistic study of the stereochemical outcomes of rhodium-catalysed styrene aziridinations. Adv. Synth. Catal 362, 384,–397.Lebel H, and Parmentier M (2010). Copper-catalyzed enantioselective aziridination of styrenes. Pure Appl. Chem 82, 1827,–1833.Lebel H, Huard K, and Lectard S (2005). N-Tosyloxycarbamates as a source of metal nitrenes: rhodium-catalyzed CH insertion and aziridination reactions. J. Am. Chem. Soc 127, 14198–14199. [Google Scholar]
  • 17.Chen Y, Fields KB, and Zhang XP (2004). Bromoporphyrins as versatile synthons for modular construction of chiral porphyrins: cobalt-catalyzed highly enantioselective and diastereoselective cyclopropanation. J. Am. Chem. Soc 126, 14718–14719. [DOI] [PubMed] [Google Scholar]
  • 18.Xu X, Lu H, Ruppel JV, Cui X, Lopez de Mesa S, Wojtas L, and Zhang XP (2011). Highly asymmetric intramolecular cyclopropanation of acceptor-substituted diazoacetates by Co(II)-based metalloradical catalysis: iterative approach for development of new-generation catalysts. J. Am. Chem. Soc 133, 15292–15295. [DOI] [PubMed] [Google Scholar]
  • 19.Heine H, and Proctor Z (1958). Notes - isomerization of N-p-Ethoxybenzolethylenimine. J. Org. Chem. 23, 1554–1556. Nishimura M, Minakata S, Takahashi T, Oderaotoshi Y, and Komatsu M (2002). Asymmetric N1 unit transfer to olefins with a chiral nitridomanganese complex: novel stereoselective pathways to aziridines or oxazolines. J. Org. Chem. 67, 2101,–2110.Luppi G, and Tomasini C (2003). A new entry to polyfunctionalized 4,5-trans disubstituted oxazolidin-2-ones from l-aspartic acid. ChemInform 34, 0797–0800. [Google Scholar]
  • 20.For previous DFT studies on Co(II)-catalyzedaziridination with sulfonyl and aryl azides using simplified model catalysts, see: Suarez AI, Jiang H, Zhang XP, and de Bruin B (2011). The radical mechanism of cobalt(II) porphyrin-catalyzed olefin aziridination and the importance of cooperative H-bonding. Dalton Trans 40, 5697–5705.Hopmann KH, and Ghosh A (2011). Mechanism of cobalt-porphyrin–catalyzed aziridination. ACS Catal 1, 597–600. [DOI] [PubMed] [Google Scholar]
  • 21.Lyaskovskyy V, Suarez AIO, Lu H, Jiang H, Zhang XP, and de Bruin B (2011). Mechanism of cobalt(II) porphyrin-catalyzed C–H amination with organic azides: radical nature and H-atom abstraction ability of the key cobalt(III)–nitrene intermediates. J. Am. Chem. Soc 133, 12264–12273. [DOI] [PubMed] [Google Scholar]
  • 22.For select examples on the significance of N–H aziridines, see 11e and:Lowden PAS (2006). Aziridine natural products – discovery, biological activity and biosynthesis. In Aziridines and Epoxides in Organic Synthesis, Yudin AK, ed. (Wiley-VCH Press; ), pp. 399–442.Lu Z, Zhang Y, and Wulff WD (2007). Direct access to N–H-aziridines from asymmetric catalytic aziridination with borate catalysts derived from vaulted binaphthol and vaulted biphenanthrol ligands. J. Am. Chem. Soc 129, 7185,–7194.Ismail FMD, Levitsky DO, and Dembitsky VM (2009). Aziridine alkaloids as potential therapeutic agents. Eur. J. Med. Chem 44, 3373,–3387.Thibodeaux CJ, Chang WC, and Liu HW (2012). Enzymatic chemistry of cyclopropane, epoxide, and aziridine biosynthesis. Chem. Rev 112, 1681–1709. [Google Scholar]
  • 23.Alonso DA, and Andersson PG (1998). Deprotection of sulfonyl aziridines. J. Org. Chem 63, 9455–9461. [Google Scholar]; Greene TW, and Wutts PG (2006). Protection for the Amino Group. In Greene’s Protective Groups in Organic Synthesis (John Wiley & Sons; ), pp. 696–926. [Google Scholar]; Ankner T, and Hilmersson G (2009). Instantaneous deprotection of tosylamides and esters with SmI2/amine/water. Org. Lett 11, 1865. [DOI] [PubMed] [Google Scholar]
  • 24.Lebel H, Lectard S, and Parmentier M (2007). Copper-catalyzed alkene aziridination with N-Tosyloxycarbamates. Org. Lett 9, 4797–4800. [DOI] [PubMed] [Google Scholar]
  • 25.Lucet D, Le Gall T, and Mioskowski C (1998). The chemistry of vicinal diamines. Angew. Chem. Int. Ed. Engl. 37, 2580–2627. [DOI] [PubMed] [Google Scholar]; Saibabu Kotti SRS, Timmons C, and Li G (2006). Vicinal diamino functionalities as privileged structural elements in biologically active compounds and exploitation of their synthetic chemistry. Chem. Biol. Drug Des 67, 101–114.Wu B, Parquette JR, and Rajanbabu TV (2009). Regiodivergent ring opening of chiral aziridines. Science 326, 1662,.Wu B, Gallucci JC, Parquette JR, and Rajanbabu TV (2009). Enantioselective Desymmetrization of meso-aziridines with TMSN3 or TMSCN Catalyzed by Discrete Yttrium Complexes. Angew. Chem. Int. Ed. Engl 48, 1126,–1129.Wu B, Gallucci JC, Parquette JR, and Rajanbabu TV (2014). Bimetallic catalysis in the highly enantioselective ring–opening reactions of aziridines. Chem. Sci 5, 1102–1117. [Google Scholar]
  • 26.For select examples on the significance ofchiral ureas, see: Breuzard JAJ, Christ-Tommasino ML, and Lemaire M (2005). Chiral ureas and Thiroureas in asymmetric catalysis. In Chiral Diazaligands for Asymmetric Synthesis, Lemaire M and Mangeney P, eds. (Springer; ), pp. 231–270. [Google Scholar]; Doyle AG, and Jacobsen EN (2007). Small-molecule H-bond donors in asymmetric catalysis. Chem. Rev 107, 5713–5743. [DOI] [PubMed] [Google Scholar]
  • 27.Hill JE, Matlock JV, Lefebvre Q, Cooper KG, and Clayden J (2018). Consecutive ring expansion and contraction for the synthesis of 1-aryl tetrahydroisoquinolines and Tetrahydrobenzazepines from readily available heterocyclic precursors. Angew. Chem. Int. Ed. Engl 57, 5788–5791. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

1
2

Data Availability Statement

The crystal structure data of compound (R)-3p, (R)-3w, and (S)-6p have been deposited in the Cambridge structural database under reference numbers CCDC: 2025803, 2025804, and 2025805, respectively.

Full experimental procedures are provided in the Supplemental information.

RESOURCES