Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Sulfenate anion catalyzed enantio- and diastereoselective aziridination

Youge Pua, Anthony M. Smaldonea, Javier Adrio*ab and Patrick J. Walsh*a
aRoy and Diana Vagelos Laboratories, Department of Chemistry, University of Pennsylvania, 231 South 34th Street, Philadelphia, PA, USA. E-mail: pwalsh@sas.upenn.edu
bDepartment of Organic Chemistry, Institute for Advance Research in Chemical Sciences (IAdChem), Universidad Autónoma de Madrid, Cantoblanco, 28049-Madrid, Spain. E-mail: javier.adrio@uam.es

Received 9th July 2025 , Accepted 29th August 2025

First published on 1st September 2025


Abstract

The synthesis of enantioenriched aziridines is important for drug development due to their prevalence in bioactive molecules. Previous methods often use expensive catalysts, activated substrates, or show poor stereoselectivity. Herein, we report a novel organocatalytic approach using enantioenriched [2.2]paracyclophane (PCP)-based sulfenate anion catalysts, enabling the synthesis of 18 cyclopropanated aziridines from unactivated imines and commercially available benzyl chlorides in 50–99% yields with 73–99% ee and >20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr. This approach fills a gap in the existing methods for aziridine synthesis, facilitating the generation of cyclopropyl-substituted aziridines with high stereoselectivity under mild and transition metal-free reaction conditions.


Introduction

Enantioenriched aziridines serve as valuable scaffolds for various biologically active natural products, such as Mitomycins1–3 and Azinomycins.4–6 These structures are known for their antitumor, antibiotic, antimicrobial, neoplasm inhibiting, and glycosidase inhibitory properties.7–9 Due to the nature of their strained rings, aziridines serve as synthetic intermediates in ring-opening reactions10,11 to produce amine-derived products.12 They are also useful in ring-expansion reactions to form larger heterocycles,13 such as β-lactams,14 γ-lactams,15 pyrrolidines,16 and piperidines,17 which are essential in both organic synthesis and medicinal chemistry.

Classic methods for the stereoselective synthesis of aziridines18–20 often involve the addition of nitrenes to olefins,21–23 transfer of carbenes to imines,24–27 and the intramolecular cyclization of chiral 1,2-vicinal haloamines28–30 or amino alcohols.31 Many of these approaches use transition metal catalysts due to their efficiency and broad applicability. However, concerns regarding the high costs and sustainability of these catalysts have driven a shift toward more environmentally friendly methods. To address this, researchers have increasingly turned to organocatalysis32,33 as a promising alternative for aziridine synthesis.

In this regard, the groups of Cordova,34,35 Hamada,36 and Albrecht37 have reported organocatalytic enantioselective aziridination reactions utilizing an aza-Michael-initiated ring-closing approach (Scheme 1a), which leads to good to excellent yields and stereoselectivities. However, these reactions, are typically limited to electronically activated substrates, such as α,β-unsaturated carbonyl compounds.


image file: d5sc05077d-s1.tif
Scheme 1 Recent advances in organocatalytic aziridination.

To expand the scope of this process, Kürti and coworkers developed a ketone catalyzed aziridination of unactivated olefins (Scheme 1b).38 This approach exploits an in situ-generated oxaziridine intermediate, enabling selective nitrogen transfer to unactivated carbon–carbon double bonds with excellent regio- and diastereoselectivity. This method has yet to be made enantioselective. Chein and coworkers39 developed an organocatalytic method for asymmetric aziridination with benzyl bromides and imines via the imino Corey–Chaykovsky reaction, using a tetrahydrothiophene-based chiral sulfide as the catalyst (Scheme 1c). This method achieved aziridination of N-phosphonate-activated benzaldimines with excellent enantioselectivities. Although these methods expand the range of accessible aziridines, the diastereoselectivities were moderate (dr = 54/46 to 83/17). Thus, complementary strategies for achieving high levels of enantio- and diastereoselectivity with unactivated substrates remain in demand.

Our group has been exploring sulfur-based organocatalysts and has successfully employed the sulfenate anion (RSO), the conjugate base of sulfenic acids, in various catalytic reactions, including the synthesis of trans-stilbenes,40,41 stilbene-based polymers42 and a one-pot method to form all three bonds of diaryl alkynes.43 The nucleophilic nature of the sulfenate anion allows it to effectively attack electrophiles such as benzyl chlorides, while its ability to function as a leaving group facilitates the closure of the catalytic cycle. The sulfur changes oxidation state in the catalytic cycle44–49 and the intermediate sulfoxide activates the α-hydrogens toward deprotonation. Higher-valent sulfur species such as S(IV)50 and S(VI)26,51–53 have also been reported, illustrating the range of accessible oxidation states. Expanding on these advances, we have recently developed a diastereoselective method for the synthesis of racemic trans-aziridines from imines and benzylic or alkyl halides using sulfenate anion (PhSO) catalysts54 (Scheme 1d). While this method affords good yields and high diastereoselectivities (trans[thin space (1/6-em)]:[thin space (1/6-em)]cis > 20[thin space (1/6-em)]:[thin space (1/6-em)]1), the catalyst is achiral, and the products are racemic.

To develop an enantioselective aziridination to couple the two electrophilic partners, we envisioned introducing an enantioenriched sulfenate anion catalyst. Herein, we report the first example of an asymmetric sulfenate anion catalyzed enantio- and diastereoselective aziridine formation (Scheme 1e) utilizing an enantioenriched [2.2]paracyclophane-substituted sulfoxide precatalyst. To our knowledge, this is the first example of asymmetric catalysis using an enantioenriched sulfenate anion.

Results and discussion

Proposed mechanism

Prior to discussing the specifics of catalyst design, we first outline the working mechanism (Fig. 1), as it informs the design process. The catalytic cycle begins with the generation of the sulfenate anion A.54 As a strong nucleophile, the sulfenate anion readily reacts with alkyl halides (B) to generate sulfoxide C with a change in oxidation state at sulfur. The resulting sulfoxide, in its more oxidized form, has activated α-hydrogens (pKa ∼ 27.2 in DMSO)55 that can be deprotonated by moderately strong bases to yield the deprotonated intermediate D. Intermediate D is also a strong nucleophile but preferentially reacts with the imine E, leading to the formation of F. The basic nitrogen in F can then act as a nucleophile, displacing the sulfenate anion and closing the catalytic cycle with formation of the aziridine G.
image file: d5sc05077d-f1.tif
Fig. 1 Proposed mechanism of enantio- and diastereoselective synthesis of aziridines.

The challenge in developing asymmetric sulfenate anion catalysts for aziridine synthesis is that there are two steps that form three new stereocenters in the catalytic reaction, as depicted in the proposed mechanism in Fig. 1. When the sulfenate anion R*SO (A) reacts with benzyl chloride, the sulfur lone pairs—previously enantiotopic in the achiral catalyst ArSO—become diastereotopic in enantioenriched R*SO. To generate a single diastereomer of the catalyst at the sulfoxide intermediate (C) the benzylation reaction must proceed with complete diastereoselectivity. We hypothesize that the configuration at sulfur in catalytic intermediate C will impact the formation of the two carbon-stereocenters in the addition adduct, D. Selecting an appropriate R* group presents a significant challenge, as the chirality of R*SO plays a role in the establishment of three contiguous stereocenters during the formation of intermediate F (one at sulfur and two at the carbons that will form the aziridine backbone).

Reaction development

We envisioned catalysts R*SO, where R* is a planar chiral para-cyclophane (PCP). Fortunately, beautiful work by the Perrio group on the stoichiometric benzylation of the rac-para-cyclophane, PCP–SO, had been reported in 2008.56 This team demonstrated that under their conditions the planar chirality of the PCP group completely controlled the central chirality at sulfur during the SN2 reaction with benzyl bromide via the depicted conformation with the stereoselectivity shown (Table 1, entry 1–3). In this conformation (A0), one sulfur lone pair extends outward from the ring, while the other is buried between the two aromatic rings. This spatial arrangement directs nucleophilic attack to occur predominantly from the more exposed lone pair, ensuring high stereoselectivity. However, our conditions for the asymmetric aziridination are expected to differ. For example, we previously demonstrated that silyl amide bases were far more effective in racemic aziridine formation than the tert-butoxide base used by the Perrio group in the benzylation.54 Furthermore, the achiral sulfenate anion catalyst employed a Li+ counterion, rather than K+ and was performed in a different solvent. Given that main group counterions and solvents are well known to significantly impact reactivity and diastereoselectivity in organic reactions,57,58 these differences are likely to have significant implications in the benzylation at sulfur.
Table 1 Diastereoselective sulfenate salt alkylation

image file: d5sc05077d-u1.tif

Entry RX Base Solvent Temp Yieldc
a Reactions performed by Perrio and coworkers.b Reaction performed using 2 equiv. of BnCl, 2 equiv. of LiN(SiMe3)2 with CPME (0.1 M).c Isolated yield.
1a BnBr tBuOK THF −78 °C 82%
2a BnBr tBuOK THF −40 °C 80%
3a BnBr tBuOK THF 0 °C 77%
4b BnCl LiN(SiMe3)2 CPME 80 °C 78%


With these considerations in mind, we chose conditions similar to our previous racemic aziridination studies,54 using LiN(SiMe3)2 to generate rac-PCP-SO at 80 °C in the presence of benzyl chloride. We were pleased to observe the formation of a single diastereomer of PCP–S([double bond, length as m-dash]O)CH2Ph (B1) in 78% yield (Table 1, entry 4). This result supports the formation of a single sulfoxide intermediate, maintaining the stereochemical integrity of the catalyst and preventing the generation of mixed diastereomers that could compromise the enantio- and diastereoselectivity in the aziridine forming steps.

For proof-of-principle studies, we resolved racemic B1 into its enantiomers using preparative chiral phase HPLC on a small scale. We then employed the enantioenriched B1 (20 mol%) as catalyst in the presence of 1 equiv. (E)-N,1-Diphenylmethanimine (1b), benzyl chloride (2a) and 2 equiv. LiN(SiMe3)2 at 80 °C with the goal of preparing enantioenriched triphenyl aziridine (Scheme 2a). Despite achieving 92% yield of the desired product 4, it was found to be racemic.


image file: d5sc05077d-s2.tif
Scheme 2 Enantioselective aziridination and studies on reversible aziridine ring-opening.

The observation of 0% ee in asymmetric catalysis is quite informative.59 It usually indicates either that the enantioenriched “catalyst” is not actually catalyzing the reaction or there is a path for rapid racemization of the enantioenriched product. To investigate the origin of the observed 0% ee, we considered the possibility of aziridine racemization via thermal ring-opening. Aziridines are known to undergo thermally induced ring-opening to form ylide intermediates under certain conditions (Scheme 2b).60,61 To assess whether such a process contributes to racemization in our system, we examined the reactivity of fluorinated triphenyl aziridine 5 with dipolarophile 6 (5 equiv.) in CPME (0.1 M) at 45 °C in the presence of an internal standard (C6F6 in C6D6). The reaction was monitored by 19F{1H} NMR spectroscopy (Scheme 2c). After 20 h at 45 °C, full conversion of the aziridine to the corresponding cycloadduct was observed (99% assay yield, determined by 19F{1H} NMR). These results suggest that under the aziridine-forming reaction conditions in Scheme 2a, the aziridine product undergoes reversible ring-opening to form the higher energy achiral azomethine ylide, leading to racemization.

We hypothesized that an alkyl substituent on the imine would destabilize the azomethine ylide, thereby increasing the energy barrier for the ring opening/racemization process. To test this hypothesis, we examined the reactivity of cyclopropanated aziridine 3a with 6 at slightly lower temperature (CPME, 60 °C; Scheme 2d). Notably, only trace amounts of the corresponding 3-pyrroline product were observed, while 96% of the starting aziridine was recovered. This experiment indicates that alkyl-substituted azomethine ylides are significantly less prone to racemization through the azomethine ylide than their aryl-substituted counterparts.

Encouraged by these findings, we next examined the reaction of imine 1b and benzyl chloride using enantioenriched B1 as the catalyst (20 mol%, Scheme 2e). Notably, the desired product was obtained in 80% yield with 87% ee, demonstrating that the presence of an alkyl group is critical for suppressing the reversible ring-opening process and preserving enantioselectivity.

Pre-catalyst design and synthesis

Next, we prepared sulfoxide pre-catalysts that enable direct entry into the catalytic cycle via the sulfenate anion.41 Based on our prior experience, we hypothesized that both pre-catalysts A1 and A2 (Scheme 3) would be synthetically accessible and capable of generating the PCP-substituted sulfenate anion via base-promoted elimination of isobutene or styrene, respectively.
image file: d5sc05077d-s3.tif
Scheme 3 Pre-catalysts for the generation of sulfenate anions through base-promoted elimination.

Both A1 (ref. 62 and 63) and A2 (ref. 64 and 65) were synthesized from rac-PCP-Br, followed by lithiation, and addition to enantioenriched sulfinyl reagents, yielding enantioenriched diastereomers that were separated by chromatography (see the SI for details). The ee of both catalysts were determined to be up to 99% by chiral stationary phase supercritical fluid chromatography (SFC).

Reaction optimization

We then employed both pre-catalysts A1 and A2 in reaction optimization studies (Table 2). We began by optimizing the reaction conditions on a 0.1 mmol scale using model substrates: (E)-1-cyclopropyl-N-(4-methoxyphenyl)methanimine (1a) and benzyl chloride (2a), with a catalyst loading of 10 mol% (unless otherwise stated). A1 was selected for initial screening due to its more straightforward and efficient synthetic preparation. In a base screening with LiN(SiMe3)2, NaN(SiMe3)2, and KN(SiMe3)2 (entries 1–3), only LiN(SiMe3)2 showed high diastereoselectivity, yielding exclusively the trans product with an assay yield (AY) of 55% (determined by 1H NMR using dibromomethane as an internal standard, Table 2). NaN(SiMe3)2, and KN(SiMe3)2 both produced a significant amount of the cis product, and no enantioselectivity was observed.
Table 2 Optimization of the aziridination reaction with (E)-1-cyclopropyl-N-(4-methoxyphenyl)methanimine (1a) and benzyl chloride (2a)

image file: d5sc05077d-u2.tif

Entry Pre-cat. Solvent Temp. Yielda,b Eec, drb
a Reaction conditions: 1a (0.10 mmol, 1.0 equiv.), 2a (0.20 mmol, 2.0 equiv.), solvent (1.0 mL, 0.1 M), 24 h.b Assay yield and dr's were determined by 1H NMR using CH2Br2 as the internal standard.c ee was determined by SFC using column Chiralcel OJ3 with 5% MeOH and 95% CO2.d NaN(SiMe3)2 was used as base.e KN(SiMe3)2 was used as base.f Pre-catalyst A1, base, and solvent were preheated at 110 °C for 30 min.g 0.2 M in CPME.h 0.4 M in CPME; 2 equiv. of 1a and 1 equiv. of 2a were used and the same yield and ee were observed.i 0.4 M in CPME and 5 mol% A2 was used.
1 A1 1,4-Dioxane 80 °C 55% 80%, >20[thin space (1/6-em)]:[thin space (1/6-em)]1
2d A1 1,4-Dioxane 80 °C 41% 0%, 3[thin space (1/6-em)]:[thin space (1/6-em)]2
3e A1 1,4-Dioxane 80 °C 76% 0%, 1[thin space (1/6-em)]:[thin space (1/6-em)]1
4f A1 1,4-Dioxane 110 °C 30 min, then 80 °C 99% 76%, >20[thin space (1/6-em)]:[thin space (1/6-em)]1
5f A1 1,4-Dioxane 110 °C 30 min, then 60 °C 99% 80%, >20[thin space (1/6-em)]:[thin space (1/6-em)]1
6f A1 1,4-Dioxane 110 °C 30 min, then 45 °C 66% 83%, >20[thin space (1/6-em)]:[thin space (1/6-em)]1
7f A1 CPME 110 °C 30 min, then 60 °C 28% 81%, >20[thin space (1/6-em)]:[thin space (1/6-em)]1
8f A1 Toluene 110 °C 30 min, then 60 °C Trace N/A
9f A1 nBu2O 110 °C 30 min, then 60 °C 16% N/A
10 A2 1,4-Dioxane 60 °C 70% 62%, >20[thin space (1/6-em)]:[thin space (1/6-em)]1
11 A2 CPME 60 °C 66% 88%, >20[thin space (1/6-em)]:[thin space (1/6-em)]1
12 A2 CPME 45 °C 70% 96%, >20[thin space (1/6-em)]:[thin space (1/6-em)]1
13 A2 CPME 25 °C 18% 94%, >20[thin space (1/6-em)]:[thin space (1/6-em)]1
14g A2 CPME 45 °C 84% 97%, >20[thin space (1/6-em)]:[thin space (1/6-em)]1
15h A2 CPME 45 °C 92% 97%, >20[thin space (1/6-em)]:[thin space (1/6-em)]1
16i A2 CPME 45 °C 76% 94%, >20[thin space (1/6-em)]:[thin space (1/6-em)]1


Recognizing that higher temperatures might be required to facilitate the elimination of isobutene from A1 and generate the sulfenate anion,41 we then started the reaction by pre-stirring the precatalyst, solvent, and base at 110 °C for 30 min. The reaction mixture was cooled to 80 °C, 60 °C, or 45 °C (entries 4–6), at which point the remaining reagents were added and the reaction was allowed to proceed for an additional 24 h. This pre-activation step significantly improved the yield from 55% (entry 1) to 99% (entries 4–5). Notably, entry 5 provided a higher ee of 80%. Further lowering the temperature did not significantly increase the ee but instead resulted in a substantial yield reduction (entry 6). Next, we evaluated a series of solvents, including CPME, toluene, and nBu2O (entries 7–9). While the reaction in CPME provided an enantioselectivity comparable to that achieved with dioxane, the yield decreased significantly. Toluene produced only trace amounts of the product (entry 8), and nBu2O resulted in a 16% yield (entry 9), both too low to obtain reliable ee and dr data. Integrating the results from base, temperature, and solvent screenings, the optimal conditions for precatalyst A1 were determined to be 10 mol% catalyst loading, 2 equiv. of LiN(SiMe3)2, and 1,4-dioxane (0.1 M), with pre-stirring at 110 °C for 30 min, followed by 60 °C for 24 h. Under these conditions, the final product 3a was obtained in 99% yield with 80% ee (entry 5). Attempts to further enhance the enantioselectivity, including the addition of various additives were unsuccessful (see the SI for details).

Given that precatalyst A2 was expected to generate the sulfenate anion more efficiently and under milder conditions, due to the greater acidity of its β-hydrogens and the stabilization of the styrene elimination product, it was subsequently used in the reaction optimization. Using the optimized conditions for pre-catalyst A1, we evaluated the two top-performing solvents with A2 as the pre-catalyst. While 1,4-dioxane achieved a slightly higher yield of 70% (entry 10), CPME provided a comparable yield of 66% (entry 11) but with a significantly improved enantiomeric excess (88% vs. 62% for 1,4-dioxane). Based on these results, CPME was selected as the solvent for further optimization. Considering that lower temperatures might enhance the enantioselectivity by favoring the pathway with the lowest activation energy, we decreased the temperature from 60 °C to 45 °C (entry 12) and then to 25 °C (entry 13). We were pleased to observe enantioselectivities of 96% and 94%, respectively. Since 45 °C maintained the yield (70%) while 25 °C resulted in only 18% yield, we selected 45 °C as the reaction temperature for further optimization. Increasing the solution concentration from 0.1 M to 0.2 M and 0.4 M resulted in a notable yield improvement (from 70% to 84% and 92%, entries 14 and 15) while holding the ee at 97%. Reducing the precatalyst loading from 10 mol% to 5 mol% led to a decrease in yield from 92% to 76% (entry 16). Adjusting the ratio of imine to benzyl chloride from 1[thin space (1/6-em)]:[thin space (1/6-em)]2 to 2[thin space (1/6-em)]:[thin space (1/6-em)]1 yielded identical results in terms of yield and ee (entry 15). However, the 2[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio provided higher yields for other substrates, likely due to reduced byproduct formation from coupling of benzyl chlorides to form trans-stilbenes.40 As a result, the optimized reaction conditions for the aziridination employed 2 equiv. LiN(SiMe3)2 in CPME (0.4 M) with the pre-catalyst loading of 10 mol% at 60 °C and with a ratio of 2[thin space (1/6-em)]:[thin space (1/6-em)]1 imine to benzyl chloride (entry 15).

Substrate scope

With the optimized conditions in hand, we investigated the substrate scope with commercially available benzyl chlorides. As shown in Scheme 4, benzyl chlorides bearing electron-donating aryl groups such as 4-Me (2b), 4-tBu (2c), 4-SMe (2d), 4-C2H4Ph (2e), and 1,3,5-Me3 (2f) provided trans-aziridines in 81–99% yield and 87–98% ee. Benzyl chlorides bearing electron-withdrawing substituents with 3-OPh (2g), 3-CF3 (2h) or 4-OCF3 (2l) cleanly afforded trans-aziridines 3g (98% yield, 97% ee), 3h (81% yield, 84% ee) and 3i (81% yield, 97% ee).
image file: d5sc05077d-s4.tif
Scheme 4 Substrate scope (dr's > 20[thin space (1/6-em)]:[thin space (1/6-em)]1). aReaction was conducted for 36 hours. bReaction was conducted with 4 equiv. of 2r.

We were particularly interested in halogen-substituted aziridines due to their potential for further functionalization via transition metal-catalyzed cross-coupling reactions. Accordingly, using benzyl chlorides substituted with 3-Br (2j) and 2-Br (2k) groups led to the formation of trans-aziridine 3j in 88% yield with 90% ee and trans-aziridine 3k in 73% yield with 73% ee. The lower yield of 3k is attributed to the steric hindrance caused by the 2-bromo substituent. For benzyl chlorides bearing 4-F or 2-F, the enantioenriched trans-aziridine 3i and 3m were obtained in 96% with >99% ee and 94% yield with 95% ee, respectively. Similarly, 4-chlorobenzyl chloride was efficiently converted to trans-aziridine 3n in 90% yield and 95% ee.

We tested 1-(chloromethyl)naphthalene, with extended conjugation, which successfully yielded trans-aziridine 3o in 70% yield and 92% ee. Carbon–carbon double bonds are also of interest due to their potential for further elaboration, such as hydrofunctionalization,66 oxidation,67 or cross-coupling reactions.68 Accordingly, 4-vinylbenzyl chloride (2p) was used to deliver trans-aziridine 3p in 93% yield with 96% ee. Aziridines containing heterocycles are frequently used in drug molecules. As shown in Scheme 4, a pyrimidine-containing aziridine 3q was obtained in 50% yield with 73% ee.

We were interested to determine if electrophiles other than benzyl chlorides were suitable. Allyl chloride was attractive because the product would be a vinyl aziridine. In the event, the use of allyl chloride (2r) furnished the vinyl-substituted aziridine 3r in 58% yield with 89% ee, highlighting the potential of this method to make highly functionalized building blocks.69–71

Other alkyl aldehydes were evaluated for imine formation, including pivaldehyde and cyclohexanecarbaldehyde, to generate (E)-N-(4-methoxyphenyl)-2,2-dimethylpropan-1-imine (2s), (E)-1-cyclohexyl-N-(4-methoxyphenyl)methanimine (2t), and (E)-1-phenyl-N-(3-(trifluoromethyl)phenyl)methanimine (2u), separately. However, imine 2s did not yield any product likely due to increased steric hindrance. For imines with α-C–H's, like 2t, tautomerization to the enamine occurred at room temperature, resulting in no formation of the desired product 3t. Unfortunately, substrates that readily form enamines are not viable under our reaction conditions. In the case of 2u, we attempted to install an electron-withdrawing 3-C6H4-CF3 group on the imine nitrogen to reduce the stability of the possible ring-opening azomethine ylide intermediate (Scheme 2b). However, the resulting aziridine 3u was obtained in 92% yield with 0% ee, indicating that racemization is still occurring under our reaction conditions due to the ring-opening process.

Gram-scale and X-ray structure determination

To illustrate the scalability of this aziridine synthesis, compound 3l was prepared on a 6 mmol scale, yielding a 95% isolated yield (1.62 g) with 99% ee and >20[thin space (1/6-em)]:[thin space (1/6-em)]1 dr (Scheme 5). To determine the absolute configuration of the aziridines, a single crystal of the aziridine 3l was obtained through cooling 3l in a solution in hexanes from 40 °C to −16 °C. The X-ray crystal structure (Fig. 2, CCDC: 2422017) confirmed that the aziridine was the (2S,3S) isomer (see the SI for crystallographic data), which was obtained by using the precatalyst A2 with a Rp configuration.
image file: d5sc05077d-s5.tif
Scheme 5 Gram-scale synthesis of aziridine 3l.

image file: d5sc05077d-f2.tif
Fig. 2 X-ray structure of (2S,3S) aziridine 3l (CCDC: 2422017).

Conclusions

In conclusion, we have developed the first enantioselective sulfenate anion-catalyzed method, which has been showcased in the highly enantioselective synthesis of trans-aziridines from simple benzyl chlorides and imines. The reaction proceeds with good to excellent yields and enantioselectivities. The [2.2]paracyclophane (PCP) scaffold is instrumental in providing excellent stereocontrol, not only in the formation of the sulfur stereocenter, but also in the generation of the two stereocenters in the aziridine backbone. Although the current scope is somewhat restricted due to ring opening of certain aziridines to achiral azomethane ylides, it provides proof-of-concept that this strategy is viable and enables the synthesis of enantioenriched cyclopropyl aziridines, a class of compounds not previously accessible with high levels of enantioselectivity. Further studies to develop more efficient oxidation state altering main group catalysts are under way in our laboratories.

Author contributions

P. J. W. conceived the project and supervised the research. J. A. conducted preliminary experiments and Y. P. conducted the experimental work and data analysis with A. S. P. J. W and Y. P. wrote the original draft of the manuscript. All authors contributed to the discussion and revision of the manuscript.

Conflicts of interest

The authors declare no competing financial interest.

Data availability

CCDC 2422017 contains the supplementary crystallographic data for this paper.72

All data for this manuscript is included in the SI. Supporting information: The experimental procedures, characterization data and crystallographic data. Deposition Number 2422017 contains the supplementary crystallographic data for this paper. See DOI: https://doi.org/10.1039/d5sc05077d.

Acknowledgements

P. J. W. thanks the NSF CHE-2154593. AMS acknowledges the NSF REU program CHE-1851640. J. A. is grateful for financial support from the Ministerio de Educación, Cultura y Deporte, subprograma estatal de movilidad and Fulbright grant visiting scholar.

Notes and references

  1. T. Hata, T. Hoshi, K. Kanamori, A. Matsumae, Y. Sano, T. Shima and R. Sugawara, Mitomycin, a new antibiotic from Streptomyces. I, J. Antibiot., Ser. A, 1956, 9, 141–146 CAS.
  2. T. Hata and R. Sugawara, Mitomycin, a new antibiotic from Streptomyces. II. Description of the strain, J. Antibiot., Ser. A, 1956, 9, 147–151 CAS.
  3. S. Wakaki, H. Marumo, K. Tomioka, G. Shimizu, E. Kato, H. Kamada, S. Kudo and Y. Fujimoto, Isolation of new fractions of antitumor mitomycins, Antibiot. Chemother., 1958, 8, 228–240 CAS.
  4. T. J. Hodgkinson and M. Shipman, Chemical synthesis and mode of action of the azinomycins, Tetrahedron, 2001, 57, 4467–4488 CrossRef CAS.
  5. G. T. Kelly, C. Liu, R. Smith, R. S. Coleman and C. M. H. Watanabe, Cellular Effects Induced by the Antitumor Agent Azinomycin B, Chem. Biol., 2006, 13, 485–492 CrossRef CAS.
  6. J. Foulke-Abel, H. Agbo, H. Zhang, S. Mori and C. M. H. Watanabe, Mode of action and biosynthesis of the azabicycle-containing natural products azinomycin and ficellomycin, Nat. Prod. Rep., 2011, 28, 693–704 RSC.
  7. P. A. S. Lowden, Aziridine Natural Products – Discovery, Biological Activity and Biosynthesis, in Aziridines and Epoxides in Organic Synthesis, ed. A. K. Yudin, 2006, pp. 399–442,  DOI:10.1002/3527607862.ch11.
  8. C. Henderson, Recent Advances in the Usage of Mitomycin. Proceedings of a symposium. Hawaii, March 21-24, 1991, Oncology, 1993, 50(suppl. 1), 1–83 Search PubMed.
  9. M. Casely-Hayford and M. Searcey, The Azinomycins. Discovery, Synthesis, and DNA-Binding Studies, in DNA and RNA Binders, ed. M. Demeunynck, C. Bailly, W. D. Wilson, 2002, pp. 676–696,  DOI:10.1002/3527601783.ch24.
  10. S. Sabir, G. Kumar, V. P. Verma and J. L. Jat, Aziridine Ring Opening: An Overview of Sustainable Methods, ChemistrySelect, 2018, 3, 3702–3711 CrossRef CAS.
  11. P. S. Bera, Y. K. Mirza, T. Sachdeva and M. Bera, Recent Advances in Transition Metal-Catalyzed Ring-Opening Reaction of Aziridine, Compounds, 2024, 4, 626–649 CrossRef CAS.
  12. X. E. Hu, Nucleophilic ring opening of aziridines, Tetrahedron, 2004, 60, 2701–2743 CrossRef CAS.
  13. N. Srivastava and H.-J. Ha, Regioselective ring opening of aziridine for synthesizing azaheterocycle, Front. Chem., 2023, 11, 1280633 CrossRef CAS.
  14. N. Piens and M. D'Hooghe, Carbonylation of Aziridines as a Powerful Tool for the Synthesis of Functionalized β-Lactams, Eur. J. Org Chem., 2017, 2017, 5943–5960 CrossRef CAS.
  15. T. B. Bisol, A. J. Bortoluzzi and M. M. Sá, Nucleophilic Ring-Opening of Epoxide and Aziridine Acetates for the Stereodivergent Synthesis of β-Hydroxy and β-Amino γ-Lactams, J. Org. Chem., 2011, 76, 948–962 CrossRef CAS.
  16. T.-Y. Lin, H.-H. Wu, J.-J. Feng and J. Zhang, Divergent Access to Functionalized Pyrrolidines and Pyrrolines via Iridium-Catalyzed Domino-Ring-Opening Cyclization of Vinyl Aziridines with β-Ketocarbonyls, Org. Lett., 2017, 19, 6526–6529 CrossRef CAS.
  17. J. Eshon, K. A. Nicastri, S. C. Schmid, W. T. Raskopf, I. A. Guzei, I. Fernández and J. M. Schomaker, Intermolecular [3+3] ring expansion of aziridines to dehydropiperi-dines through the intermediacy of aziridinium ylides, Nat. Commun., 2020, 11, 1273 CrossRef CAS PubMed.
  18. L. Degennaro, P. Trinchera and R. Luisi, Recent Advances in the Stereoselective Synthesis of Aziridines, Chem. Rev., 2014, 114, 7881–7929 CrossRef CAS PubMed.
  19. D. Tanner, Chiral aziridines—their synthesis and use in stereoselective transformations, Angew Chem. Int. Ed. Engl., 1994, 33, 599–619 CrossRef.
  20. H. M. I. Osborn and J. Sweeney, The asymmetric synthesis of aziridines, Tetrahedron Asymmetry, 1997, 8, 1693–1715 CrossRef CAS.
  21. L. Maestre, W. M. C. Sameera, M. M. Díaz-Requejo, F. Maseras and P. J. Pérez, A General Mechanism for the Copper- and Silver-Catalyzed Olefin Aziridination Reactions: Concomitant Involvement of the Singlet and Triplet Pathways, J. Am. Chem. Soc., 2013, 135, 1338–1348 CrossRef CAS.
  22. Y. Li, J. He, V. Khankhoje, E. Herdtweck, K. Köhler, O. Storcheva, M. Cokoja and F. E. Kühn, Copper(ii) complexes incorporating poly/perfluorinated alkoxyaluminate-type weakly coordinating anions: syntheses, characterization and catalytic application in stereoselective olefin aziridination, Dalton Trans., 2011, 40, 5746–5754 RSC.
  23. D. A. Evans, M. T. Bilodeau and M. M. Faul, Development of the Copper-Catalyzed Olefin Aziridination Reaction, J. Am. Chem. Soc., 1994, 116, 2742–2753 CrossRef CAS.
  24. K. B. Hansen, N. S. Finney and E. N. Jacobsen, Carbenoid Transfer to Imines: A New Asymmetric Catalytic Synthesis of Aziridines, Angew Chem. Int. Ed. Engl., 1995, 34, 676–678 CrossRef CAS.
  25. J. R. Krumper, M. Gerisch, J. M. Suh, R. G. Bergman and T. D. Tilley, Monomeric Rhodium(II) Catalysts for the Preparation of Aziridines and Enantioselective Formation of Cyclopropanes from Ethyl Diazoacetate at Room Temperature, J. Org. Chem., 2003, 68, 9705–9710 CrossRef CAS.
  26. X. Zeng, X. Zeng, Z. Xu, M. Lu and G. Zhong, Highly Efficient Asymmetric Trans-Selective Aziridination of Diazoacetamides and N-Boc-imines Catalyzed by Chiral Brønsted Acids, Org. Lett., 2009, 11, 3036–3039 CrossRef CAS.
  27. A.-H. Li, L.-X. Dai and V. K. Aggarwal, Asymmetric Ylide Reactions: Epoxidation, Cyclopropanation, Aziridination, Olefination, and Rearrangement, Chem. Rev., 1997, 97, 2341–2372 CrossRef CAS PubMed.
  28. D. Guijarro, Ó. Pablo and M. Yus, Synthesis of γ-, δ-, and ε-Lactams by Asymmetric Transfer Hydrogenation of N-(tert-Butylsulfinyl)iminoesters, J. Org. Chem., 2013, 78, 3647–3654 CrossRef CAS.
  29. Y. Hayashi, T. Urushima, D. Sakamoto, K. Torii and H. Ishikawa, One-Pot Synthesis of Chiral Aziridines by a Domino Reaction by Using Desulfonylative Formation on the N-Tosyl Imine of Chloroacetaldehyde with an Asymmetric Mannich Reaction as a Key Step, Chem.–Eur. J., 2011, 17, 11715–11718 CrossRef CAS.
  30. J. Wei, Z. Chen, Y. Gao, P. Zhang, C. Wang, P. Zhao, Y. Wang and X. Shi, A Rapid and Simple Method for Quantitative Aziridination from Aminobrominated Derivatives of Olefins under Solvent-free and Mild Conditions, Chin. J. Chem., 2012, 30, 391–399 CrossRef CAS.
  31. S. K. Kim and E. N. Jacobsen, General Catalytic Synthesis of Highly Enantiomerically Enriched Terminal Aziridines from Racemic Epoxides, Angew. Chem., Int. Ed., 2004, 43, 3952–3954 CrossRef CAS.
  32. E. Roma, E. Tosi, M. Miceli and T. Gasperi, Asymmetric Organocatalytic Aziridination: Recent Advances, Asian J. Org. Chem., 2018, 7, 2357–2367 CrossRef CAS.
  33. Y. Zhu, Q. Wang, R. G. Cornwall and Y. Shi, Organocatalytic Asymmetric Epoxidation and Aziridination of Olefins and Their Synthetic Applications, Chem. Rev., 2014, 114, 8199–8256 CrossRef CAS.
  34. L. Deiana, G.-L. Zhao, S. Lin, P. Dziedzic, Q. Zhang, H. Leijonmarck and A. Córdova, Organocatalytic Enantioselective Aziridination of α-Substituted α,β-Unsaturated Aldehydes: Asymmetric Synthesis of Terminal Aziridines, Adv. Synth. Catal., 2010, 352, 3201–3207 CrossRef CAS.
  35. J. Vesely, I. Ibrahem, G.-L. Zhao, R. Rios and A. Córdova, Organocatalytic Enantioselective Aziridination of α,β-Unsaturated Aldehydes, Angew. Chem., Int. Ed., 2007, 46, 778–781 CrossRef CAS.
  36. H. Arai, N. Sugaya, N. Sasaki, K. Makino, S. Lectard and Y. Hamada, Enantioselective aziridination reaction of α,β-unsaturated aldehydes using an organocatalyst and tert-butyl N-arenesulfonyloxycarbamates, Tetrahedron Lett., 2009, 50, 3329–3332 CrossRef CAS.
  37. S. Frankowski, J. Bojanowski, M. Saktura, M. Romaniszyn, P. Drelich and Ł. Albrecht, Organocatalytic Synthesis of cis-2,3-Aziridine Aldehydes by a Postreaction Isomerization, Org. Lett., 2017, 19, 5000–5003 CrossRef CAS PubMed.
  38. Q.-Q. Cheng, Z. Zhou, H. Jiang, J. H. Siitonen, D. H. Ess, X. Zhang and L. Kürti, Organocatalytic nitrogen transfer to unactivated olefins via transient oxaziridines, Nat. Catal., 2020, 3, 386–392 CrossRef CAS.
  39. M.-T. Huang, H.-Y. Wu and R.-J. Chein, Enantioselective synthesis of diaryl aziridines using tetrahydrothiophene-based chiral sulfides as organocatalysts, Chem. Commun., 2014, 50, 1101–1103 RSC.
  40. M. Zhang, T. Jia, H. Yin, P. J. Carroll, E. J. Schelter and P. J. Walsh, A New Class of Organocatalysts: Sulfenate Anions, Angew. Chem., Int. Ed., 2014, 53, 10755–10758 CrossRef CAS PubMed.
  41. M. Zhang, T. Jia, I. K. Sagamanova, M. A. Pericás and P. J. Walsh, tert-Butyl Phenyl Sulfoxide: A Traceless Sulfenate Anion Precatalyst, Org. Lett., 2015, 17, 1164–1167 CrossRef CAS.
  42. M. Li, S. Berritt, C. Wang, X. Yang, Y. Liu, S.-C. Sha, B. Wang, R. Wang, X. Gao, Z. Li, X. Fan, Y. Tao and P. J. Walsh, Sulfenate anions as organocatalysts for benzylic chloromethyl coupling polymerization via C=C bond formation, Nat. Commun., 2018, 9, 1754 CrossRef PubMed.
  43. M. Zhang, T. Jia, C. Y. Wang and P. J. Walsh, Organocatalytic Synthesis of Alkynes, J. Am. Chem. Soc., 2015, 137, 10346–10350 CrossRef CAS.
  44. D. M. Freudendahl, S. Santoro, S. A. Shahzad, C. Santi and T. Wirth, Green Chemistry with Selenium Reagents: Development of Efficient Catalytic Reactions, Angew. Chem., Int. Ed., 2009, 48, 8409–8411 CrossRef CAS.
  45. A. Breder and S. Ortgies, Recent developments in sulfur- and selenium-catalyzed oxidative and isohypsic functionalization reactions of alkenes, Tetrahedron Lett., 2015, 56, 2843–2852 CrossRef CAS.
  46. S. Ortgies and A. Breder, Oxidative Alkene Functionalizations via Selenium-π-Acid Catalysis, ACS Catal., 2017, 7, 5828–5840 CrossRef CAS.
  47. L. Shao, Y. Li, J. Lu and X. Jiang, Recent progress in selenium-catalyzed organic reactions, Org. Chem. Front., 2019, 6, 2999–3041 RSC.
  48. F. V. Singh and T. Wirth, Selenium reagents as catalysts, Catal. Sci. Technol., 2019, 9, 1073–1091 RSC.
  49. A. Matviitsuk, J. L. Panger and S. E. Denmark, Catalytic, Enantioselective Sulfenofunctionalization of Alkenes: Development and Recent Advances, Angew. Chem., Int. Ed., 2020, 59, 19796–19819 CrossRef CAS PubMed.
  50. X. Zhang, R. Sun, D. Zeng, M. Wang and X. Jiang, Cu-catalyzed Bilateral Linkage of Sulfenamides for the Construction of Benzothiophene-based Cyclic Sulfilimines, Eur. J. Org Chem., 2025, 28, e202401467 CrossRef CAS.
  51. M. Wang, S. Chen and X. Jiang, Construction of Functionalized Annulated Sulfone via SO2/I Exchange of Cyclic Diaryliodonium Salts, Org. Lett., 2017, 19, 4916–4919 CrossRef CAS.
  52. D. Zeng, Y. Ma, W.-P. Deng, M. Wang and X. Jiang, Divergent sulfur(VI) fluoride exchange linkage of sulfonimidoyl fluorides and alkynes, Nat. Synth., 2022, 1, 455–463 CrossRef CAS.
  53. S. Zhao, D. Zeng, M. Wang and X. Jiang, C-SuFEx linkage of sulfonimidoyl fluorides and organotrifluoroborates, Nat. Commun., 2024, 15, 727 CrossRef CAS.
  54. Z. Zheng, Y. Pu, J. Adrio and P. J. Walsh, Sulfenate Anion Catalyzed Diastereoselective Synthesis of Aziridines, Angew. Chem., Int. Ed., 2023, 62, e202303069 CrossRef CAS.
  55. F. G. Bordwell, Equilibrium acidities in dimethyl sulfoxide solution, Acc. Chem. Res., 1988, 21, 456–463 CrossRef CAS.
  56. J.-F. Lohier, F. Foucoin, P.-A. Jaffrès, J. I. Garcia, J. Sopková−de Oliveira Santos, S. Perrio and P. Metzner, An Efficient and Straightforward Access to Sulfur Substituted [2.2]Paracyclophanes: Application to Stereoselective Sulfenate Salt Alkylation, Org. Lett., 2008, 10, 1271–1274 CrossRef CAS PubMed.
  57. T. X. Gentner and R. E. Mulvey, Alkali-Metal Mediation: Diversity of Applications in Main-Group Organometallic Chemistry, Angew. Chem., Int. Ed., 2021, 60, 9247–9262 CrossRef CAS PubMed.
  58. R. Sreedharan and T. Gandhi, Masters of Mediation: MN(SiMe3)2 in Functionalization of C(sp)−H Latent Nucleophiles, Chem.–Eur. J., 2024, 30, e202400435 CrossRef CAS.
  59. P. J. Walsh and M. C. Kozlowski, Fundamentals of Asymmetric Catalysis, University Science Books, 2009 Search PubMed.
  60. I. Coldham and R. Hufton, Intramolecular Dipolar Cycloaddition Reactions of Azomethine Ylides, Chem. Rev., 2005, 105, 2765–2810 CrossRef CAS.
  61. H. W. Heine and R. Peavy, Aziridines X.I. Reaction of 1,2,3-triphenylaziridine with diethylacetylene dicarboxylate and maleic anhydride, Tetrahedron Lett., 1965, 6, 3123–3126 CrossRef.
  62. G. J. Rowlands and R. J. Seacome, Enantiospecific synthesis of [2.2]paracyclophane-4-thiol and derivatives, Beilstein J. Org. Chem., 2009, 5, 9 Search PubMed.
  63. D. J. Weix and J. A. Ellman, Improved Synthesis of tert-Butanesulfinamide Suitable for Large-Scale Production, Org. Lett., 2003, 5, 1317–1320 CrossRef CAS.
  64. P. B. Hitchcock, G. J. Rowlands and R. Parmar, The synthesis of enantiomerically pure 4-substituted [2.2]paracyclophane derivatives by sulfoxide–metal exchange, Chem. Commun., 2005, 4219–4221,  10.1039/B507394D.
  65. D. Jishkariani, B. T. Diroll, M. Cargnello, D. R. Klein, L. A. Hough, C. B. Murray and B. Donnio, Dendron-Mediated Engineering of Interparticle Separation and Self-Assembly in Dendronized Gold Nanoparticles Superlattices, J. Am. Chem. Soc., 2015, 137, 10728–10734 CrossRef CAS.
  66. X. Yin, S. Li, K. Guo, L. Song and X. Wang, Palladium-Catalyzed Enantioselective Hydrofunctionalization of Alkenes: Recent Advances, Eur. J. Org Chem., 2023, 26, e202300783 CrossRef CAS.
  67. C. Bonini and G. Righi, A critical outlook and comparison of enantioselective oxidation methodologies of olefins, Tetrahedron, 2002, 58, 4981–5021 CrossRef CAS.
  68. L. J. Oxtoby, J. A. Gurak Jr, S. R. Wisniewski, M. D. Eastgate and K. M. Engle, Palladium-Catalyzed Reductive Heck Coupling of Alkenes, Trends Chem., 2019, 1, 572–587 CrossRef CAS PubMed.
  69. D. Kang, T. Kim, H. Lee and S. Hong, Regiodivergent Ring-Opening Cross-Coupling of Vinyl Aziridines with Phosphorus Nucleophiles: Access to Phosphorus-Containing Amino Acid Derivatives, Org. Lett., 2018, 20, 7571–7575 CrossRef CAS.
  70. D. C. D. Butler, G. A. Inman and H. Alper, Room Temperature Ring-Opening Cyclization Reactions of 2-Vinylaziridines with Isocyanates, Carbodiimides, and Isothiocyanates Catalyzed by [Pd(OAc)2]/PPh3, J. Org. Chem., 2000, 65, 5887–5890 CrossRef CAS PubMed.
  71. T. Zhang, S. Wang, D. Zuo, J. Zhao, W. Luo, C. Wang and P. Li, Palladium-Catalyzed Carbonylative [5+1] Cycloaddition of N-Tosyl Vinylaziridines: Solvent-Controlled Divergent Synthesis of α,β- and β,γ-Unsaturated δ-Lactams, J. Org. Chem., 2022, 87, 10408–10415 CrossRef CAS PubMed.
  72. Y. Pu, A. M. Smaldone, J. Adrio, P. Walsh, CCDC, Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2m99lv.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.