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

Monodentate halogen bond activation of aziridines in formal [3 + 2] cycloadditions

Mattis Damrath and Boris J. Nachtsheim*
Institute for Organic and Analytical Chemistry, University of Bremen, 28359 Bremen, Germany. E-mail: nachtsheim@uni-bremen.de

Received 5th August 2025 , Accepted 16th September 2025

First published on 17th September 2025


Abstract

The ring opening of aziridines to produce various N-heterocycles traditionally requires strong Lewis acids or transition metal catalysts, with non-covalent organocatalytic approaches remaining largely unexplored. Herein, we demonstrate that N-heterocyclic iodonium salts can effectively catalyze [3 + 2] cycloadditions of aziridines through a monodentate halogen bond (XB) activation. Using 1–5 mol% of the iodolium catalyst, a wide range of aziridines undergo an efficient cycloaddition with a variety of dipolarophiles (carbonyls, alkynes, and alkenes) to furnish oxazolidines, pyrrolines, and pyrrolidines. DFT calculations revealed a previously underexplored N-activation mode, with detailed non-covalent interaction analysis showing that the N-heterocyclic iodonium salt's exceptional performance stems from combined I–N and I–π interactions.


Introduction

Non-covalent interactions have been demonstrated to play a hitherto unparalleled role in the supramolecular assembly of proteins,1–4 in enzymatic reactions,5–7 as well as in drug design and material science.8–15 Among the extensive array of non-covalent interactions, hydrogen bond donors have shown notable efficacy in their role as organocatalysts, facilitating a diverse spectrum of reactions.16–21 In recent years, halogen bond (XB)22,23 and chalcogen bond (ChB)24 donors have emerged as significant contributors to the field of organocatalysis.25–32 Also, these compounds have been recognized for their substantial contributions in crystal engineering,33,34 molecular recognition,35,36 and in medicinal applications.37,38 The restriction of reactivity imposed by non-covalent interactions delineates the constraints of this catalysis discipline, particularly regarding the limitations of activation targets and reaction patterns. Despite the capacity of halogen or chalcogen bond donors to activate a variety of molecules, including carbonyl compounds,39–42 imines,43 and nitroolefins,44–46 their application has recently been extended to more challenging substrates such as esters,47 alkenes,48 alkynes,49 and even carbene precursors.50,51 Nevertheless, the activation of strained heterocycles such as aziridines remains a formidable challenge, as these transformations have historically relied on strong Lewis acids, transition metals, or high temperatures (Scheme 1A).52–55 In 2022, Wang and co-workers presented the first example of a bidentate selenium(II) catalyst for the non-covalent activation of aziridines in a [3 + 2] cycloaddition (Scheme 1B).56 They demonstrated the ability to obtain different N-heterocycles by altering the dipolarophile from alkenes to alkynes and carbonyls in good yields. Further NMR experiments demonstrated that a cooperative Se–O and Se–N activation, facilitating the bidentate mode, in conjunction with the precise distance between the two selenium units, was instrumental in achieving the desired outcome. In a recent study, Tan and co-workers introduced a novel class of bidentate organotellurium catalysts, which demonstrated remarkable efficacy in the activation of various azetidines and aziridines for the cycloaddition with alkenes and alkynes.57
image file: d5qo01119a-s1.tif
Scheme 1 (A) Typical approach for the activation of aziridines by strong Lewis acids, transition metals, or elevated temperatures. (B) First example of non-covalent organocatalysis for the ring opening of aziridines by a bidentate chalcogen bond catalyst by Wang and co-workers. (C) This work: XB-catalyzed [3 + 2] cycloaddition of aziridines with carbonyls, alkynes, and alkenes.

Given the well-established fact that XB-donors facilitate significantly more robust activation than their chalcogen analogs, our research group was intrigued by the question of whether hypercoordinated iodine compounds,58–64 especially our previously developed N-heterocyclic iodonium salt XB catalysts,65–67 would also allow an effective activation of aziridines for [3 + 2]-cycloadditions.

In the present work, we demonstrate that such an activation by a monodentate XB catalyst is a straightforward process that allows the synthesis of a variety of oxazolidine, pyrroline, and pyrrolidine scaffolds from N-sulfonyl aziridines.

Results and discussion

Reaction development

Initially, the reactivity of this activation process by halogen bonding was evaluated through the investigation of the well-established monodentate XB catalysts 1a–e in the ring opening of N-tosyl-protected aziridine 2a with cyclohexanone (3a) to obtain oxazolidine 4a (Scheme 2). Initial preoptimizations with catalyst 1a (see SI) revealed that a catalyst loading of only 1 mol% of the XB-donor in DCM was sufficient to successfully synthesize the target compound 4a in a nearly quantitative yield after 2 h.
image file: d5qo01119a-s2.tif
Scheme 2 1H-NMR kinetic studies for the cycloaddition of aziridine 2a with cyclohexanone (3a) employing various iodine(III) XB-donors 1a–e. General reaction conditions: 2a (0.100 mmol, 1.0 eq.), 3a (3.0 eq.), XB catalyst 1 (1 mol%), DCM (0.1 M), rt.

After implementing the aforementioned conditions, a 1H-NMR kinetic study was conducted with all well-established catalysts 1a–e.42,65–67 This kinetic study revealed that catalyst 1a exhibited the highest level of activity, surpassing all the other iodonium salt catalysts. It should be noted that the N-heterocyclic iodonium salt catalysts 1b and 1d have shown comparable conversion rates to each other and provided good yields of about 80% after 4 h reaction time. It is also noteworthy that the C-bound pyrazole 1c exhibited a significant decrease in activity, with a yield of only 22% after the specified time, while using the simple iodolium salt catalyst 1e resulted in almost negligible product formation.

The observed effectiveness of these catalysts in activating aziridines aligns with the relative Lewis acidity strength of the tested monodentate iodonium salts, with higher acidity generally producing better results.68 This observation suggests that the efficiency of the halogen bond-catalyzed reaction depends heavily on the strength of the catalyst's Lewis acid properties. To further verify the relevance of the XB for the catalytic activation, the reaction was carried out with catalyst 1a in the presence of 10 mol% NBu4Br. This experiment resulted in no product formation, indicating an XB-mediated activation of the aziridine 2a.

After establishing the aforementioned conditions, other substrates were investigated. Through time-dependent 1H-NMR spectroscopy, we could prove that when utilizing benzaldehyde (3b) as a coupling partner, a complete conversion to oxazolidine 4b could be obtained after a reaction time of approximately 30 min (Scheme 3A).


image file: d5qo01119a-s3.tif
Scheme 3 XB-catalyzed cycloaddition between aziridine 2a and benzaldehyde (3b) with 1a as the XB donor. General reaction conditions: 2a (0.100 mmol, 1.0 equiv.), 3b (3.0 equiv.), 1a (1 mol%), DCM (0.1 M), rt. (A) Kinetic investigation of diastereoselectivity. (B) Kinetic investigation with the addition of NBu4Br (10 mol%) after 15 min.

This reaction exhibited a diastereomeric ratio (dr) of 92[thin space (1/6-em)]:[thin space (1/6-em)]8, in favor of the cis-isomer (cis-4b).69 We also observed that as the reaction continued, the diastereomeric ratio underwent a shift. After a reaction time of 1 h, a commendable dr of 90[thin space (1/6-em)]:[thin space (1/6-em)]10 was obtained; however, after 2 h, a ratio of 78[thin space (1/6-em)]:[thin space (1/6-em)]22 was already observed, which resulted after 4 h in a final dr of 45[thin space (1/6-em)]:[thin space (1/6-em)]55, with the trans-4b isomer being slightly favored. The experiment provided clear evidence that the initial kinetic preference is gradually overridden as the catalyst continues to mediate reversible ring-opening/ring-closing sequences, ultimately approaching thermodynamic equilibrium where the slightly more stable trans-4b isomer marginally predominates. To further verify the observed behavior, we repeated the reaction but added NBu4Br (10 mol%) after an initial reaction time of 15 min (Scheme 3B). Analysis of the 1H-NMR spectra over a 3-hour period clearly showed that addition of the Lewis base completely halted the reaction. This confirms our hypothesis that catalyst 1a causes the observed time-dependent conversion between diastereomers. It also highlights the importance of continuous reaction monitoring to properly determine the scope of suitable substrates.

Scope of the reaction with carbonyls

With this information in hand, the reaction of aziridine 2a with benzaldehyde (3b) was terminated after 30 min with an excellent yield of 93% and 92[thin space (1/6-em)]:[thin space (1/6-em)]8 dr for the target oxazolidine 4b (Scheme 4). With sterically hindered 2-methylbenzaldehyde (3c), we observed a significant decrease in the conversion rate. After 24 h, oxazolidine 4c could still be obtained in 85% yield with an improved diastereoselectivity of >95[thin space (1/6-em)]:[thin space (1/6-em)]5. The cycloaddition of 2a with the electron-donating 4-methoxybenzaldehyde (3d) yielded the desired oxazolidine 4d after 30 min in nearly quantitative yield of 96% but with no diastereoselectivity. The electron-withdrawing nitro-substituted oxazolidine 4e was also accessible under the investigated reaction conditions with 90% yield and 95[thin space (1/6-em)]:[thin space (1/6-em)]5 dr after 2 h. Next, we focused on aromatic carbonyl compounds. Acetophenone, indanone, and benzophenone proved to be suitable coupling partners for the cycloaddition of aziridine 2a, with yields of up to 97% and up to 75[thin space (1/6-em)]:[thin space (1/6-em)]25 dr for the oxazolidines 4f–4h. We also conducted the cycloaddition with acetone, which resulted in a comparable and nearly quantitative yield of 99% for 4i. Using N-methylisatin (3j) as the substrate, the reaction time increased to 24 h with a moderate yield of 62% and 3.5[thin space (1/6-em)]:[thin space (1/6-em)]1 dr for the desired spirocycle 4j.
image file: d5qo01119a-s4.tif
Scheme 4 Substrate scope for the XB-catalyzed cycloaddition of aziridines 2 and carbonyls 3 toward oxazolidines 4. General reaction conditions: 2 (0.100 mmol, 1.0 equiv.), 3 (3.0 equiv.), 1a (1 mol%), DCM (0.1 M), rt. a[thin space (1/6-em)]Reaction time: 30 min. b[thin space (1/6-em)]Reaction time: 24 h. c[thin space (1/6-em)]Reaction time: 6 h.

Next, a variety of aziridines were investigated under our reaction conditions with cyclohexanone as the coupling reagent. First, various protecting groups on the aziridine moiety were compared, and it was found that both the Nosyl (Ns) and Mesyl (Ms) protecting groups could be used, although significantly longer reaction times were required compared to the Tosyl protecting group, and lower yields for 4k and 4l of 64% and 91% were observed. Other protecting groups, such as Acetyl or Boc, did not show any reactivity under the reaction conditions investigated. Concerning the substitution on the aromatic skeleton of the aziridine, electron-rich, electron-poor or slightly sterically demanding derivatives 4o–s were accessible in yields of up to 96%.

Aliphatic aziridines, such as 2t and 2u, could also be converted via XB-catalysis in the cycloaddition, but yielded oxazolidines 4t and 4u in significantly lower yields of 44% and 73%. Herein, we observed the formation of several by-products, including the ring-opening of the oxazolidine, which are likely attributable to the extended reaction times of 6 h and 24 h. The indene-derived aziridine 2v could again be converted to oxazolidine 4v in a short reaction time with a high yield of 95%, which indicates the requirement of the aromatic functionality of the aziridine for a fast and effective reaction outcome. The more activated aziridines 4w and 4x were unsuitable substrates. In both cases, known Lewis acid-mediated rearrangements occurred (see SI). Even adjusting the reaction conditions, such as lower temperatures or an increased catalyst loading, did not allow a successful reaction. To gain deeper insights into this reaction's regioselectivity, we conducted the cycloaddition of aziridine 2y. We were able to separate the two resulting regioisomers 4y and 4z in 31% and 19% yield, respectively, with the higher substituted one as the expected major regioisomer.

Notably, our reaction protocol also worked well for opening the four-membered ring compound azetidine 5. When combined with benzaldehyde, this reaction produced oxazinane 6 with good stereoselectivity (single diastereomer) and yield (75%).

Scope of the reaction with alkynes

Beyond creating oxazolidines 4 from aziridines 2 and carbonyls 3, we found that our established reaction conditions also effectively promoted the cycloaddition between aziridines and alkynes 7 to form pyrrolines 8 (Scheme 5). Indeed, the reaction with phenylacetylene with 2a afforded the desired pyrroline 8a in an excellent yield of 91%. To further extend the range of aryl alkynes, various substituents on the aryl skeleton were investigated under the reaction conditions. While slightly sterically-hindered or electron-donating alkynes could be efficiently converted to the desired pyrrolines 8b–d in yields of up to 92%, the use of electron-deficient alkynes afforded only traces of the target compounds 8e and 8f. Furthermore, diphenylacetylene and 1-phenylpropyne were identified as suitable coupling reagents, yielding the corresponding pyrrolines 8g and 8h in 79% and 88% yield, respectively. The limitations of the method could be demonstrated through the utilization of alkyl or TMS-substituted derivatives, which are less activated or sterically hindered. Using these alkyne derivatives to potentially give pyrrolines 8i–k resulted either in complete absence of product formation or yielded inseparable mixtures of compounds, demonstrating the limitations of this methodology with sterically hindered or electronically deactivated substrates. In the context of the aziridine scope, the evaluation of various aromatic substituents on the aziridine skeleton was conducted. The employment of the slightly sterically-hindered 2-methyl-substituted aziridine proved to be efficient within the reaction conditions, and the product 8l could be obtained in a yield of 89%.
image file: d5qo01119a-s5.tif
Scheme 5 Substrate scope for the XB-catalyzed cycloaddition of aziridines 2 and alkynes 7 toward pyrrolines 8. General reaction conditions: 2 (0.100 mmol, 1.0 equiv.), 7 (3.0 equiv.), 1a (1 mol%), DCM (0.1 M), rt. a[thin space (1/6-em)]1a (5 mol%) was used. b[thin space (1/6-em)]Reaction time: 30 min. c[thin space (1/6-em)]Reaction time: 24 h. d[thin space (1/6-em)]Reaction time: 1 h. e[thin space (1/6-em)]Reaction temperature: 50 °C.

The formation of the highly activated pyrroline 8m was unsuccessful due to product decomposition in the purification process. Pyrrolines 8n and 8o, containing OAc- or Br-substitutions, were obtained in yields of 85% and 88%. In contrast, the electron-deficient NO2-substituted derivative could only be accomplished in a yield of 29%. After an extended reaction time of 24 h. The efficacy of alternative protective groups was also demonstrated, exhibiting a reaction behavior analogous to that observed with carbonyl derivatives. The Ns-protected pyrroline 8q and the Ms-protected derivative 8r could be isolated in yields of 47% and 88%. The pyrroline 8s was accessible under the developed reaction conditions, but only in a moderate yield of 50%. Our method showed clear limitations when applied to aliphatic aziridines like the cyclohexene-derived compound, as we detected no formation of the target pyrroline 8t. Similar to our earlier findings with carbonyl compounds, we found that azetidine 5 (the four-membered ring analog) could react with phenylacetylene to produce tetrahydropyridine 9, though with only 35% yield.

Scope of the reaction with alkenes

Inspired by the study conducted by Wang and co-workers,56 we also investigated the ring opening of aziridines with alkene compounds to obtain pyrrolidines. Initial experiments conducted under the previously utilized reaction conditions with aziridine 2a and styrene (10a) yielded low yields of the desired compound, accompanied by long reaction times and the formation of multiple side products. Increasing the catalyst loading to 5 mol% effectively addressed the aforementioned issues, thereby enabling the exploration of a substrate scope for the cycloaddition of aziridines with alkene derivatives. Unlike the cycloadditions with carbonyl compounds, we found that diastereoselectivity could not be controlled in these reactions. 1H-NMR monitoring revealed that both diastereomers formed in nearly equal amounts right at the beginning of the reaction. For example, the initial reaction of aziridine 2a with styrene (10a) afforded a promising yield of 87% and 1[thin space (1/6-em)]:[thin space (1/6-em)]1 dr for the desired pyrrolidine 11a after 1 h reaction time (Scheme 6).
image file: d5qo01119a-s6.tif
Scheme 6 Substrate scope for the XB-catalyzed cycloaddition of aziridines 2 and alkenes 10 toward pyrrolidines 11. General reaction conditions: 2 (0.100 mmol, 1.0 equiv.), 10 (3.0 equiv.), 1a (5 mol%), DCM (0.1 M), rt. a[thin space (1/6-em)]Reaction time: 24 h. b[thin space (1/6-em)]Reaction time: 2 h. c[thin space (1/6-em)]Reaction temperature: 50 °C. d[thin space (1/6-em)]Reaction time: 4 h. e[thin space (1/6-em)]Reaction time: 8 h.

In the following, we started our investigation by examining various aromatic substituents on the styrene derivative to induce a potential diastereoselectivity characterized by electronic or steric properties. Using the sterically-hindered 2-methylstyrene, pyrrolidine 11b was obtained in 63% yield with 1.3[thin space (1/6-em)]:[thin space (1/6-em)]1 dr, indicating a negligible impact of steric substituents on diastereoselectivity. Furthermore, para-alkyl substituted pyrrolidines 11c,d were generated in high yields of up to 90%, while the strong electron-donating methoxy substituted derivative 11e was unsuitable due to its rapid polymerization. Other para substituents such as Ph, OAc or Br afforded the desired products 11f–h in up to 72% and up to 1.2[thin space (1/6-em)]:[thin space (1/6-em)]1 dr. The strong electron-deficient pyrrolidine 11i could only be detected in traces after an extended reaction time of 24 h. Subsequently, the focus shifted to the investigation of substitutions on the exocyclic double bond of styrene. For trisubstituted pyrrolidines 11j–l, yields ranged from 42% to 91%. In case of indene as the dipolarophile, an inseparable mixture of the product 11m and the non-cyclized derivative was obtained (see SI). As an example of alkyl substitution, the reaction with isobutene yielded dimethylpyrrolidine 11n in 85% yield. It was also possible to convert 1,3-dimethylindole as an alkene source under the developed conditions, leading to hexahydropyrrolo[2,3-b]indole 11o in 37% yield and with 1[thin space (1/6-em)]:[thin space (1/6-em)]1 dr, whose core structure plays a relevant role in biology and pharmacology.70

Next, we focused on the conversion of various aziridines. Both electron-donating and electron-withdrawing aromatic substituents have shown good performance with yields ranging from 68% to 87% and up to 1.6[thin space (1/6-em)]:[thin space (1/6-em)]1 dr for pyrrolidines 11p–11t, indicating a minor influence on the diastereoselectivity by these substituents. Regarding the protecting groups, the Nosyl-protected pyrrolidine 11u was accessible in a moderate yield of 58%, while the Mesyl-protected derivative 11v was unstable during the purification process. Trisubstituted pyrrolidine 11w was synthesized in 62% yield with 4[thin space (1/6-em)]:[thin space (1/6-em)]1 dr, whereas the cyclohexene-derived aziridine showed no conversion toward 11x. As with the other coupling reagents, styrene could also be converted with azetidine 5 in a formal cycloaddition to piperidine 12 in a yield of 61% and 1.8[thin space (1/6-em)]:[thin space (1/6-em)]1 dr.

Computational investigation

To gain insight into the coordination of the N-heterocyclic iodolium catalyst 1a (N- vs. O-coordination) to the substrate and the origin of the high activity of this catalyst, DFT calculations were performed on the halogen-bond catalyzed formal cycloaddition between aziridine 2a and acetone (Fig. 1). Free energies obtained at the ωB97X-D3BJ/def-2 TZVP (SMD18, DCM)//ωB97X-D3BJ/def-2 SVP (SMD18, DCM) level of theory at 298 K were used for the discussion (see SI for further details). The coordination of the catalyst 1a2+ to the aziridine 2a results in the formation of two complexes, N-I and O-I, that were found to be energetically very close, with a slight preference for the O-coordination of 0.6 kcal mol−1 (Fig. 1A).
image file: d5qo01119a-f1.tif
Fig. 1 DFT computations of the key intermediates for O- and N-activation at the ωB97X-D3BJ/def-2 TZVP (SMD18, DCM)//ωB97X-D3BJ/def-2 SVP (SMD18, DCM) level of theory at 298 K. (A) Free energy profile (in kcal mol−1) for the cycloaddition between aziridine 2a and acetone to oxazolidine 4i. (B) Non-covalent interaction analysis of the key transition states. Color code for NCI plots: red = repulsion, green = van der Waals attraction/weak interaction (e.g., hydrogen bonding, halogen bonding), blue = strong interaction (e.g., hydrogen bonding, halogen bonding).

Following the incorporation of acetone into the activation complex, the disparity between the two intermediates, N-II and O-II, diminishes to 0.2 kcal mol−1. The XB-catalyzed bond-breaking process then occurs via N-TS1 (19.6 kcal mol−1) or O-TS1 (25.5 kcal mol−1) to yield the dipole intermediates N-III (1.4 kcal mol−1) and O-III (13.2 kcal mol−1). This substantial discrepancy in the transition state energies prompted further investigation into the behavior, which was subsequently analyzed in detail using the non-covalent interaction analysis (NCIPlot) method (Fig. 1B). This analysis clearly demonstrates that O-TS1 interacts through an I–O interaction (2.38 Å), as well as a pronounced π–π interaction between the aromatic part of the iodonium salt and the aromatic unit of the aziridine. In contrast, N-TS1 exhibits, in addition to these interactions, a further I–π interaction (3.43 Å) with the second σ-hole of the catalyst toward the aromatic skeleton of the aziridine. This interaction, in combination with the presumably better stability between the negative charge on the nitrogen and the positive one on the catalyst (ion pair), clearly favors the latter transition state (ΔΔG = 5.9 kcal mol−1). This hypothesis is further substantiated by the resulting dipole intermediates, where the observed difference increased to 11.8 kcal mol−1. A comparison with iodolium salt 1e, which is experimentally inactive, reveals endergonic activation energies and significantly higher transition states (see SI for further discussion), yet the interaction behavior is the same for both pathways.

Subsequent rotations to N-IV and O-IV enable the following ring closure step. In the case of O-coordination, this intermediate can be converted barrierless into the stable catalyst-product complex O-V (−7.8 kcal mol−1). In the case of the N-coordinated reaction pathway, the barrier for bond formation is predominantly attributable to the increased stability of the dipole intermediate due to the previously mentioned ion pair. This bond formation transition state N-TS2 with a free energy of 18.0 kcal mol−1, subsequently results in a stable catalyst-product complex N-V (1.0 kcal mol−1). Decomplexation yields the oxazolidine 4i (−5.3 kcal mol−1) and regenerates the catalyst 1a2+. In summary, the favored N-coordination pathway demonstrates a rate-determining step (RDS) barrier of 19.6 kcal mol−1 and is exergonic by 5.3 kcal mol−1.

Conclusions

In conclusion, a straightforward method for the synthesis of various five-membered N-heterocycles from aziridines via XB-catalyzed [3 + 2] cycloadditions through monodentate activation is presented. N-Sulfonyl-protected aziridines have been shown to efficiently react in a formal [3 + 2] cycloaddition with carbonyls, alkynes, or alkenes as dipolarophiles in high yields and partially controllable diastereoselectivity. The exceptional performance of the N-heterocyclic iodonium salt catalyst as the XB-donor was ascertained through DFT calculations of the reaction pathway and an examination of the non-covalent interactions in the bond-breaking process. The monodentate N-coordination pathway, which has emerged as the favored reaction mode for this specific catalyst type, is still underexplored and presents opportunities for addressing additional activation targets through XB-catalysis in the future.

Methods

General procedure for cycloaddition of aziridines with various dipolarophiles

Under an argon atmosphere, a dry 10 mL tube (screw cap with septum) was charged with 1a (1.6 mg, 1.00 μmol, 1 mol% or 8.2 mg, 5.00 μmol, 5 mol% in case of alkenes as substrates) and aziridine 2 (0.100 mmol, 1.0 equiv.), evacuated three times, and backfilled with argon. Subsequently, DCM (1.0 mL) and the corresponding carbonyl 3 or alkyne 7 or alkene 10 (0.300 mmol, 3.0 equiv.) were added to the reaction mixture. The reaction was stirred until completion of the reaction (1H-NMR analysis), nBu4NBr (3.2 mg, 10.0 μmol, 10 mol%) was added, and the solvent was evaporated. The residue was purified by column chromatography to give the desired N-heterocycles.

Computational details

All structures were preoptimized using a combination of XTB71 and CREST72,73 to evaluate the lowest energy conformers. The optimization and frequency analysis of these conformers were calculated using ORCA 6.0.174–78 with the ωB97X-D3BJ/def2-SVP78–83 level of theory at room temperature (298 K). This combination was selected due to the complex binding nature of the σ-hole interactions, and its performance has been demonstrated with analogous systems.84–87 The solvent effects, utilizing DCM as the solvent, were incorporated into the optimization through the implemented solvation model based on the density approach (SMD)88 in ORCA 6.0.1 in its slightly modified version (SMD18)89 with a corrected radius for iodine atoms. Single point energies were performed on the lowest energy small basis set calculations using ωb97X-D3BJ/def2-TZVP. Relative free energies (ΔG) were calculated as the energy difference between the complexes or the transition states (TS) and the sum of the energies of the reactants and the catalyst. Transition states were verified by the intrinsic reaction coordinate (IRC)90 method as implemented in ORCA 6.0.1. The reaction energy diagram was initially created using EveRplot91 and then modified with ChemDraw. The structures were visualized with CYLView92 and the hydrogen atoms were omitted for clarity. The analysis of non-covalent interactions (NCI) was performed with NCIPlot93–95 and visualized by VMD.96 The coordinates of all calculated structures are given in the SI.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this study, including detailed experimental procedures, spectroscopic data, and additional information, are available in the supplementary information (SI). Supplementary information: experimental procedures and the analytical data, including the corresponding NMR-spectra for unknown compounds and the Cartesian coordinates of the optimized structures from the DFT calculations. See DOI: https://doi.org/10.1039/d5qo01119a.

References

  1. R. S. Spolar, J. H. Ha and M. T. Record, Hydrophobic effect in protein folding and other noncovalent processes involving proteins, Proc. Natl. Acad. Sci. U. S. A., 1989, 86, 8382–8385 CrossRef CAS PubMed.
  2. T. Young, R. Abel, B. Kim, B. J. Berne and R. A. Friesner, Motifs for molecular recognition exploiting hydrophobic enclosure in protein–ligand binding, Proc. Natl. Acad. Sci. U. S. A., 2007, 104, 808–813 CrossRef CAS PubMed.
  3. S. Jena, J. Dutta, K. D. Tulsiyan, A. K. Sahu, S. S. Choudhury and H. S. Biswal, Noncovalent interactions in proteins and nucleic acids: beyond hydrogen bonding and π-stacking, Chem. Soc. Rev., 2022, 51, 4261–4286 RSC.
  4. V. A. Adhav and K. Saikrishnan, The Realm of Unconventional Noncovalent Interactions in Proteins: Their Significance in Structure and Function, ACS Omega, 2023, 8, 22268–22284 CrossRef CAS PubMed.
  5. D. H. Williams, E. Stephens, D. P. O'Brien and M. Zhou, Understanding Noncovalent Interactions: Ligand Binding Energy and Catalytic Efficiency from Ligand-Induced Reductions in Motion within Receptors and Enzymes, Angew. Chem., Int. Ed., 2004, 43, 6596–6616 CrossRef CAS PubMed.
  6. H. He, W. Tan, J. Guo, M. Yi, A. N. Shy and B. Xu, Enzymatic Noncovalent Synthesis, Chem. Rev., 2020, 120, 9994–10078 CrossRef CAS PubMed.
  7. M. Raynal, P. Ballester, A. Vidal-Ferran and P. W. N. M. van Leeuwen, Supramolecular catalysis. Part 2: artificial enzyme mimics, Chem. Soc. Rev., 2014, 43, 1734–1787 RSC.
  8. C. C. J. Loh, Exploiting non-covalent interactions in selective carbohydrate synthesis, Nat. Rev. Chem., 2021, 5, 792–815 CrossRef CAS PubMed.
  9. A. S. Mahadevi and G. N. Sastry, Cooperativity in Noncovalent Interactions, Chem. Rev., 2016, 116, 2775–2825 CrossRef CAS PubMed.
  10. A. J. Neel, M. J. Hilton, M. S. Sigman and F. D. Toste, Exploiting non-covalent π interactions for catalyst design, Nature, 2017, 543, 637–646 CrossRef CAS PubMed.
  11. A. Haque, K. M. Alenezi, M. S. Khan, W.-Y. Wong and P. R. Raithby, Non-covalent interactions (NCIs) in π-conjugated functional materials: advances and perspectives, Chem. Soc. Rev., 2023, 52, 454–472 RSC.
  12. J. Chen, Q. Peng, X. Peng, H. Zhang and H. Zeng, Probing and Manipulating Noncovalent Interactions in Functional Polymeric Systems, Chem. Rev., 2022, 122, 14594–14678 CrossRef CAS PubMed.
  13. L. Brunsveld, B. J. B. Folmer, E. W. Meijer and R. P. Sijbesma, Supramolecular Polymers, Chem. Rev., 2001, 101, 4071–4098 CrossRef CAS PubMed.
  14. G. R. Desiraju, Crystal Engineering: From Molecule to Crystal, J. Am. Chem. Soc., 2013, 135, 9952–9967 CrossRef CAS PubMed.
  15. M. Raynal, P. Ballester, A. Vidal-Ferran and P. W. N. M. van Leeuwen, Supramolecular catalysis. Part 1: non-covalent interactions as a tool for building and modifying homogeneous catalysts, Chem. Soc. Rev., 2014, 43, 1660–1733 RSC.
  16. P. R. Schreiner, Metal-free organocatalysis through explicit hydrogen bonding interactions, Chem. Soc. Rev., 2003, 32, 289–296 RSC.
  17. M. S. Taylor and E. N. Jacobsen, Asymmetric Catalysis by Chiral Hydrogen-Bond Donors, Angew. Chem., Int. Ed., 2006, 45, 1520–1543 CrossRef CAS PubMed.
  18. A. G. Doyle and E. N. Jacobsen, Small-Molecule H-Bond Donors in Asymmetric Catalysis, Chem. Rev., 2007, 107, 5713–5743 CrossRef CAS PubMed.
  19. X. Yu and W. Wang, Hydrogen-Bond-Mediated Asymmetric Catalysis, Chem. – Asian J., 2008, 3, 516–532 CrossRef CAS PubMed.
  20. M. Terada, Chiral Phosphoric Acids as Versatile Catalysts for Enantioselective Transformations, Synthesis, 2010, 1929–1982 CrossRef CAS.
  21. R. Maji, S. C. Mallojjala and S. E. Wheeler, Chiral phosphoric acid catalysis: from numbers to insights, Chem. Soc. Rev., 2018, 47, 1142–1158 RSC.
  22. P. Metrangolo and G. Resnati, Halogen bonding: fundamentals and applications, Springer, 2008 Search PubMed.
  23. G. Cavallo, P. Metrangolo, R. Milani, T. Pilati, A. Priimagi, G. Resnati and G. Terraneo, The Halogen Bond, Chem. Rev., 2016, 116, 2478–2601 CrossRef CAS PubMed.
  24. L. Vogel, P. Wonner and S. M. Huber, Chalcogen Bonding: An Overview, Angew. Chem., Int. Ed., 2019, 58, 1880–1891 CrossRef CAS PubMed.
  25. D. Bulfield and S. M. Huber, Halogen Bonding in Organic Synthesis and Organocatalysis, Chem. – Eur. J., 2016, 22, 14434–14450 CrossRef CAS PubMed.
  26. R. L. Sutar and S. M. Huber, Catalysis of Organic Reactions through Halogen Bonding, ACS Catal., 2019, 9, 9622–9639 CrossRef CAS.
  27. J. Bamberger, F. Ostler and O. G. Mancheño, Frontiers in Halogen and Chalcogen-Bond Donor Organocatalysis, ChemCatChem, 2019, 11, 5198–5211 CrossRef CAS PubMed.
  28. S. Huber, Halogen Bonding in Solution, John Wiley & Sons, 2020 Search PubMed.
  29. M. Breugst and J. J. Koenig, σ-Hole Interactions in Catalysis, Eur. J. Org. Chem., 2020, 5473–5487 CrossRef CAS.
  30. G. Sekar, V. V. Nair and J. Zhu, Chalcogen bonding catalysis, Chem. Soc. Rev., 2024, 53, 586–605 RSC.
  31. P. W. van Leeuwen and M. Raynal, Supramolecular catalysis: new directions and developments, John Wiley & Sons, 2021 Search PubMed.
  32. D. Jovanovic, M. Poliyodath Mohanan and S. M. Huber, Halogen, Chalcogen, Pnictogen, and Tetrel Bonding in Non-Covalent Organocatalysis: An Update, Angew. Chem., Int. Ed., 2024, 63, e202404823 CrossRef CAS PubMed.
  33. P. Metrangolo, F. Meyer, T. Pilati, G. Resnati and G. Terraneo, Halogen Bonding in Supramolecular Chemistry, Angew. Chem., Int. Ed., 2008, 47, 6114–6127 CrossRef CAS PubMed.
  34. L. C. Gilday, S. W. Robinson, T. A. Barendt, M. J. Langton, B. R. Mullaney and P. D. Beer, Halogen Bonding in Supramolecular Chemistry, Chem. Rev., 2015, 115, 7118–7195 CrossRef CAS PubMed.
  35. M. S. Taylor, Anion recognition based on halogen, chalcogen, pnictogen and tetrel bonding, Coord. Chem. Rev., 2020, 413, 213270 CrossRef CAS.
  36. R. Kampes, S. Zechel, M. D. Hager and U. S. Schubert, Halogen bonding in polymer science: towards new smart materials, Chem. Sci., 2021, 12, 9275–9286 RSC.
  37. R. Wilcken, M. O. Zimmermann, A. Lange, A. C. Joerger and F. M. Boeckler, Principles and Applications of Halogen Bonding in Medicinal Chemistry and Chemical Biology, J. Med. Chem., 2013, 56, 1363–1388 CrossRef CAS PubMed.
  38. N. K. Shinada, A. G. de Brevern and P. Schmidtke, Halogens in Protein–Ligand Binding Mechanism: A Structural Perspective, J. Med. Chem., 2019, 62, 9341–9356 CrossRef CAS PubMed.
  39. S. H. Jungbauer, S. M. Walter, S. Schindler, L. Rout, F. Kniep and S. M. Huber, Activation of a carbonyl compound by halogen bonding, Chem. Commun., 2014, 50, 6281–6284 RSC.
  40. J.-P. Gliese, S. H. Jungbauer and S. M. Huber, A halogen-bonding-catalyzed Michael addition reaction, Chem. Commun., 2017, 53, 12052–12055 RSC.
  41. A. Dreger, P. Wonner, E. Engelage, S. M. Walter, R. Stoll and S. M. Huber, A halogen-bonding-catalysed Nazarov cyclisation reaction, Chem. Commun., 2019, 55, 8262–8265 RSC.
  42. M. Damrath, A. Döring and B. J. Nachtsheim, Halogen bond-catalyzed Pictet–Spengler reaction, Chem. Commun., 2025, 61, 4828–4831 RSC.
  43. M. V. Il'in, Y. V. Safinskaya, D. A. Polonnikov, A. S. Novikov and D. S. Bolotin, Chalcogen- and Halogen-Bond-Donating Cyanoborohydrides Provide Imine Hydrogenation, J. Org. Chem., 2024, 89, 2916–2925 CrossRef PubMed.
  44. P. Wonner, A. Dreger, L. Vogel, E. Engelage and S. M. Huber, Chalcogen Bonding Catalysis of a Nitro-Michael Reaction, Angew. Chem., Int. Ed., 2019, 58, 16923–16927 CrossRef CAS PubMed.
  45. C. J. Vonnemann, E. Engelage, J. S. Ward, K. Rissanen, S. Keller and S. M. Huber, Ortho-Carborane-Derived Halogen Bonding Organocatalysts, Angew. Chem., Int. Ed., 2025, e202424072,  DOI:10.1002/anie.202424072.
  46. F. Heinen, D. L. Reinhard, E. Engelage and S. M. Huber, A Bidentate Iodine(III)-Based Halogen-Bond Donor as a Powerful Organocatalyst, Angew. Chem., Int. Ed., 2021, 60, 5069–5073 CrossRef CAS PubMed.
  47. Z. Zhao, Y. Liu and Y. Wang, Weak Interaction Activates Esters: Reconciling Catalytic Activity and Turnover Contradiction by Tailored Chalcogen Bonding, J. Am. Chem. Soc., 2024, 146, 13296–13305 CrossRef CAS PubMed.
  48. X. Yuan and Y. Wang, A Selenide Catalyst for the Activation of Alkenes through Se⋯π Bonding, Angew. Chem., Int. Ed., 2022, 61, e202203671 CrossRef CAS PubMed.
  49. Y. Pang, Z. Zhao and Y. Wang, Activation of alkynes by chalcogen bonding: a Se⋯π interaction catalyzed intramolecular cyclization of 1,6-diynes, Chem. Commun., 2023, 59, 12278–12281 RSC.
  50. H. Zhu, X. Yang, Y. Liu, H. Zhou and Y. Wang, Chalcogen Bonding Catalysis Enables Ring-Opening of Cyclopropene and Ring Expansion of Aryl Ketones, Angew. Chem., Int. Ed., 2025, 64, e202423746 CrossRef CAS PubMed.
  51. Z. Zhao and Y. Wang, Chalcogen Bonding Catalysis with Phosphonium Chalcogenide (PCH), Acc. Chem. Res., 2023, 56, 608–621 CrossRef CAS PubMed.
  52. A. L. Cardoso and T. M. V. D. Pinho e Melo, Aziridines in Formal [3 + 2] Cycloadditions: Synthesis of Five-Membered Heterocycles, Eur. J. Org. Chem., 2012, 6479–6501 CrossRef CAS.
  53. X. E. Hu, Nucleophilic ring opening of aziridines, Tetrahedron, 2004, 60, 2701–2743 CrossRef CAS.
  54. P. Lu, Recent developments in regioselective ring opening of aziridines, Tetrahedron, 2010, 66, 2549–2560 CrossRef CAS.
  55. J.-J. Feng and J. Zhang, Synthesis of Unsaturated N-Heterocycles by Cycloadditions of Aziridines and Alkynes, ACS Catal., 2016, 6, 6651–6661 CrossRef CAS.
  56. H. Zhu, P.-P. Zhou and Y. Wang, Cooperative chalcogen bonding interactions in confined sites activate aziridines, Nat. Commun., 2022, 13, 3563 CrossRef CAS PubMed.
  57. F. Jiang, L.-Z. Zuo, H.-W. Zhao, S.-H. Xiang and B. Tan, Rational Design of an Organotellurium Chalcogen Bonding Catalyst for Azetidine Activation, Angew. Chem., Int. Ed., 2025, e202511650 Search PubMed.
  58. A. Yoshimura and V. V. Zhdankin, Advances in Synthetic Applications of Hypervalent Iodine Compounds, Chem. Rev., 2016, 116, 3328–3435 CrossRef CAS PubMed.
  59. A. Yoshimura and V. V. Zhdankin, Recent Progress in Synthetic Applications of Hypervalent Iodine(III) Reagents, Chem. Rev., 2024, 124, 11108–11186 CrossRef CAS PubMed.
  60. F. V. Singh, S. E. Shetgaonkar, M. Krishnan and T. Wirth, Progress in organocatalysis with hypervalent iodine catalysts, Chem. Soc. Rev., 2022, 51, 8102–8139 RSC.
  61. T. Wirth, Hypervalent Iodine Chemistry, Springer, 2016 Search PubMed.
  62. R. Robidas, D. L. Reinhard, C. Y. Legault and S. M. Huber, Iodine(III)-Based Halogen Bond Donors: Properties and Applications, Chem. Rec., 2021, 21, 1912–1927 CrossRef CAS PubMed.
  63. X. Peng, A. Rahim, W. Peng, F. Jiang, Z. Gu and S. Wen, Recent Progress in Cyclic Aryliodonium Chemistry: Syntheses and Applications, Chem. Rev., 2023, 123, 1364–1416 CrossRef CAS PubMed.
  64. D. Jovanovic and S. M. Huber, Cyclic Diaryliodonium Species as Synthetic Precursors and Halogen Bonding Organocatalysts, Chem. Rev., 2023, 123, 10527–10529 CrossRef CAS PubMed.
  65. A. Boelke, T. J. Kuczmera, L. D. Caspers, E. Lork and B. J. Nachtsheim, Iodolopyrazolium Salts: Synthesis, Derivatizations, and Applications, Org. Lett., 2020, 22, 7261–7266 CrossRef CAS PubMed.
  66. A. Boelke, T. J. Kuczmera, E. Lork and B. J. Nachtsheim, N-Heterocyclic Iod(az)olium Salts – Potent Halogen-Bond Donors in Organocatalysis, Chem. – Eur. J., 2021, 27, 13128–13134 CrossRef CAS PubMed.
  67. E. M. Galathri, T. J. Kuczmera, B. J. Nachtsheim and C. G. Kokotos, Organocatalytic Friedel–Crafts arylation of aldehydes with indoles utilizing N-heterocyclic iod(az)olium salts as halogen-bonding catalysts, Green Chem., 2024, 26, 825–831 RSC.
  68. J. Su, Y. Liu, Y. Jing, Y. Liu and Z. Ke, Theoretical Quantification of the Lewis Acidity of N-Heterocyclic Iodonium Salts, Asian J. Org. Chem., 2023, 12, e202300210 CrossRef CAS.
  69. S. Teranishi, K. Maeda, T. Kurahashi and S. Matsubara, Diastereoselective Synthesis of 1,3-Oxazolidines via Cationic Iron Porphyrin-catalyzed Cycloaddition of Aziridines with Aldehydes, Org. Lett., 2019, 21, 2593–2596 CrossRef CAS PubMed.
  70. D. Crich and A. Banerjee, Chemistry of the Hexahydropyrrolo[2,3-b]indoles: Configuration, Conformation, Reactivity, and Applications in Synthesis, Acc. Chem. Res., 2007, 40, 151–161 CrossRef CAS PubMed.
  71. C. Bannwarth, E. Caldeweyher, S. Ehlert, A. Hansen, P. Pracht, J. Seibert, S. Spicher and S. Grimme, Extended tight-binding quantum chemistry methods, Wiley Interdiscip. Rev.:Comput. Mol. Sci., 2021, 11, e1493 CAS.
  72. P. Pracht, F. Bohle and S. Grimme, Automated exploration of the low-energy chemical space with fast quantum chemical methods, Phys. Chem. Chem. Phys., 2020, 22, 7169–7192 RSC.
  73. S. Grimme, Exploration of Chemical Compound, Conformer, and Reaction Space with Meta-Dynamics Simulations Based on Tight-Binding Quantum Chemical Calculations, J. Chem. Theory Comput., 2019, 15, 2847–2862 CrossRef CAS PubMed.
  74. F. Neese, The ORCA program system, Wiley Interdiscip. Rev.:Comput. Mol. Sci., 2012, 2, 73–78 CAS.
  75. F. Neese, Software update: the ORCA program system, version 4.0, Wiley Interdiscip. Rev.:Comput. Mol. Sci., 2018, 8, e1327 Search PubMed.
  76. F. Neese, Software update: The ORCA program system—Version 5.0, Wiley Interdiscip. Rev.:Comput. Mol. Sci., 2022, 12, e1606 Search PubMed.
  77. F. Neese, F. Wennmohs, U. Becker and C. Riplinger, The ORCA quantum chemistry program package, J. Chem. Phys., 2020, 152, 224108 CrossRef CAS PubMed.
  78. A. Najibi and L. Goerigk, The Nonlocal Kernel in van der Waals Density Functionals as an Additive Correction: An Extensive Analysis with Special Emphasis on the B97M-V and ωB97M-V Approaches, J. Chem. Theory Comput., 2018, 14, 5725–5738 CrossRef CAS.
  79. J.-D. Chai and M. Head-Gordon, Long-range corrected hybrid density functionals with damped atom–atom dispersion corrections, Phys. Chem. Chem. Phys., 2008, 10, 6615–6620 RSC.
  80. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.
  81. S. Grimme, S. Ehrlich and L. Goerigk, Effect of the damping function in dispersion corrected density functional theory, J. Comput. Chem., 2011, 32, 1456–1465 CrossRef CAS PubMed.
  82. F. Weigend and R. Ahlrichs, Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy, Phys. Chem. Chem. Phys., 2005, 7, 3297–3305 RSC.
  83. F. Weigend, Accurate Coulomb-fitting basis sets for H to Rn, Phys. Chem. Chem. Phys., 2006, 8, 1057–1065 RSC.
  84. N. Melnyk, M. R. Garcia and C. Trujillo, Halogen-Bond-Based Organocatalysis Unveiled: Computational Design and Mechanistic Insights, ACS Catal., 2023, 13, 15505–15515 CrossRef CAS.
  85. J. O'Brien, N. Melnyk, R. S. Lee, M. James and C. Trujillo, Computational Design of Bidentate Hypervalent Iodine Catalysts in Halogen Bond-Mediated Organocatalysis, ChemPhysChem, 2024, 25, e202400515 CrossRef PubMed.
  86. M. H. Kolář and P. Hobza, Computer Modeling of Halogen Bonds and Other σ-Hole Interactions, Chem. Rev., 2016, 116, 5155–5187 CrossRef PubMed.
  87. S. Kozuch and J. M. L. Martin, Halogen Bonds: Benchmarks and Theoretical Analysis, J. Chem. Theory Comput., 2013, 9, 1918–1931 CrossRef CAS.
  88. A. V. Marenich, C. J. Cramer and D. G. Truhlar, Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions, J. Phys. Chem. B, 2009, 113, 6378–6396 CrossRef CAS PubMed.
  89. E. Engelage, N. Schulz, F. Heinen, S. M. Huber, D. G. Truhlar and C. J. Cramer, Refined SMD Parameters for Bromine and Iodine Accurately Model Halogen-Bonding Interactions in Solution, Chem. – Eur. J., 2018, 24, 15983–15987 CrossRef CAS PubMed.
  90. K. Ishida, K. Morokuma and A. Komornicki, The intrinsic reaction coordinate. An ab initio calculation for HNC→HCN and H−+CH4→CH4+H−, J. Chem. Phys., 1977, 66, 2153–2156 CrossRef CAS.
  91. M. K. Bogdos and B. Morandi, EveRplot: A Web-Based Shiny Application for Creating Energy vs Reaction Coordinate Diagrams, J. Chem. Educ., 2023, 100, 3641–3644 CrossRef CAS.
  92. C. Y. Legault, 1.0b, Université de Sherbrooke, CYLview, 2009, https://www.cylview.org, (accessed 2025–08–04).
  93. R. A. Boto, F. Peccati, R. Laplaza, C. Quan, A. Carbone, J.-P. Piquemal, Y. Maday and J. Contreras-García, NCIPLOT4: Fast, Robust, and Quantitative Analysis of Noncovalent Interactions, J. Chem. Theory Comput., 2020, 16, 4150–4158 CrossRef CAS PubMed.
  94. J. Contreras-García, E. R. Johnson, S. Keinan, R. Chaudret, J.-P. Piquemal, D. N. Beratan and W. Yang, NCIPLOT: A Program for Plotting Noncovalent Interaction Regions, J. Chem. Theory Comput., 2011, 7, 625–632 CrossRef PubMed.
  95. E. R. Johnson, S. Keinan, P. Mori-Sánchez, J. Contreras-García, A. J. Cohen and W. Yang, Revealing Noncovalent Interactions, J. Am. Chem. Soc., 2010, 132, 6498–6506 CrossRef CAS PubMed.
  96. W. Humphrey, A. Dalke and K. Schulten, VMD - Visual Molecular Dynamics, J. Mol. Graphics, 1996, 14, 33–38 CrossRef CAS PubMed.

This journal is © the Partner Organisations 2025
Click here to see how this site uses Cookies. View our privacy policy here.