banner
News center
We are happy to fulfill your customized request

Chalcogen bond

Apr 09, 2024

Nature Communications volume 13, Article number: 4793 (2022) Cite this article

6188 Accesses

7 Citations

6 Altmetric

Metrics details

Conformational isomerization can be guided by weak interactions such as chalcogen bonding (ChB) interactions. Here we report a catalytic strategy for asymmetric access to chiral sulfoxides by employing conformational isomerization and chalcogen bonding interactions. The reaction involves a sulfoxide bearing two aldehyde moieties as the substrate that, according to structural analysis and DFT calculations, exists as a racemic mixture due to the presence of an intramolecular chalcogen bond. This chalcogen bond formed between aldehyde (oxygen atom) and sulfoxide (sulfur atom), induces a conformational locking effect, thus making the symmetric sulfoxide as a racemate. In the presence of N–heterocyclic carbene (NHC) as catalyst, the aldehyde moiety activated by the chalcogen bond selectively reacts with an alcohol to afford the corresponding chiral sulfoxide products with excellent optical purities. This reaction involves a dynamic kinetic resolution (DKR) process enabled by conformational locking and facile isomerization by chalcogen bonding interactions.

Non–covalent interactions based on hydrogen bond1,2,3 and halogen bond4,5,6,7 represent a powerful and promising activation mode in catalytic synthesis. However, the chalcogen bond is a new class of weak non–covalent interactions between the chalcogen atom (S, Se, Te) and Lewis base (Fig. 1a), which attracted attentions only in recent years8,9,10. In the living systems, the chalcogen bonding interactions play a crucial role in regulating protein conformations11 and preserving certain enzymatic activities12,13 (Fig. 1b). These interactions have also been studied in the areas of solid–state chemistry14, anion recognition15,16,17, supramolecular assembling18,19,20, and drug designs21,22. For example, the conformational locking effect induced by chalcogen bonds is believed to enhance the bioactivities of multiple commercial pharmaceuticals such as Acetazolamide23 and Selenazofurin24. (Fig. 1b). In contrast to the relatively wide applications in functional molecule design, chalcogen bonds are much less explored as effective tools for catalysis and organic synthesis especially in asymmetrical reactions25. The use of chalcogen bonding (ChB) for catalysis received reasonable attentions only in recent years26,27. As disclosed by Matile28,29, Huber30,31 and Wang32,33,34, the key is to install chalcogen bond donors to the catalysts that can interact with the substrate for catalytic activations (Fig. 1c). Most of the success for effective catalysis comes from cationic chalcogen bonding interactions, which cationic charges are introduced to decrease the electron density of chalcogen atom to enhance chalcogen bonding interaction. Despite these impressive progresses, the development of effective chalcogen bonding catalysis remains slow, and evidences for the presence of chalcogen bond in catalytic reactions mostly relies on in situ NMR spectra (13C, 77Se)27,32,33,34,35, UV-vis and nanoESI-MS15 analysis. We postulate that part of the reasons lie on the difficulties in designing these stable chalcogen bonded complex between catalysts and substrates.

a Chalcogen bond (ChB). b ChB in living systems, medicines, and agrochemicals. c Intermolecular (cationic) ChB in organic catalysis. d Intramolecular (neutral) ChB of substrate as enabling tools for asymmetric synthesis (of chiral sulfoxides). e Examples of functional chiral sulfoxides.

We’re particularly motivated by the fact that such intramolecular interactions are widely present (or can be readily installed) in both macro18,19,20 and small molecules of natural origins11,12,13 or chemical synthesis25,26,27,28,29,30,31,32,33,34,36. It is also encouraging to observe that chalcogen bonding in the intramolecular fashion can be designed in readily predictable and modular manners21,22,23,24. For instance, Tomada et al reported a chiral selenenylation reagent bearing the intramolecular N-Se interaction to rigidify the whole molecule37. Subsequently, Wirth extended this concept towards O-Se interactions and achieved the asymmetric functionalization of alkenes38. Furthermore, the transient intramolecular chalcogen bonding interactions had been proven by Smith et al which were the crucial force in stereoselectivity control25,39,40,41,42,43,44. Based on these thought-provoking applications of intramolecular ChBs45,46,47,48, our interests are directed toward employing intramolecular chalcogen bonding interactions for conformational regulations and selective chemical transformations.

In this work, we study the involvement of ChB interactions in the reactivity of sulfinyldibenzaldehyde compounds. These noncovalent interactions are key to achieve the regioselective mono-esterification of the compound via chiral carbene-catalyzed oxidation process, enabling the preparation of chiral sulfoxides. These results indicate that ChB interactions can play an important role in asymmetric organic synthesis.

Here we disclosed a catalytic dynamic kinetic resolution protocol of sulfoxides which were enabled by intramolecular chalcogen bond–guided conformational isomerization of the substrate (Fig. 1d). The sulfoxide group have been testified to be a ChB donor8,49 and the solid–state X–ray structure of 1a (Fig. 1d, details see Supplementary Table 2) suggests the presence of chalcogen bond between sulfur and oxygen atom (bond length = 2.861 Å). This chalcogen bond breaks the symmetry of 1a, and therefore makes this symmetric sulfoxide present as a racemic mixture with two conformational enantiomers [(Rp)–1a and (Sp)–1a, Fig. 1d]. Interconversion of the two enantiomers via chalcogen bond–guided conformational isomerization was estimated by DFT, which was a facile process with an activation energy of 8.55 kcal/mol. We postulated that the chalcogen bonding likely exist in solution as well, and then started to screen suitable conditions for a dynamic kinetic resolution of such conformers by using an NHC–catalyzed esterification process. Under the catalysis of N–heterocyclic carbene (NHC)50,51,52,53,54 at an oxidative condition to covert one of the aldehyde moieties of 1a to an ester, a highly efficient dynamic kinetic resolution of this sulfoxide (1a) is realized. Oxidation of Breslow intermediates to the corresponding acyl azolium intermediates (I and I’) were estimated (via DFT) to be the stereoselectivity–determining step in this DKR process. Our reaction affords chiral sulfoxide products with good yields and excellent enantiomeric purities. Notably, chiral sulfoxides are widely used in medicines (such as Esomeprazole55 and Armodafinil56), agrochemicals (such as Ethiprole57), and as ligands in asymmetric catalysis58,59 (Fig. 1e). The chiral sulfoxides from our reactions may work as platform scaffolds for transforming to bioactive molecules and catalysts.

At first, we chose conformational isomeric sulfinyldibenzaldehyde 1a as the model sulfoxide substrate and methanol 2a as a nucleophile to search for suitable conditions, and the key results were summarized in Table 1. Triazoliums were explored as the NHC pre–catalysts with diphenoquinone (DQ)60 as an oxidant to convert one of the aldehyde moieties of 1a to an ester unit. An encouraging result was obtained when aminoindanol–derived triazoium A was the NHC pre–catalyst with K2CO3 as a base in THF, offering the corresponding chiral sulfoxide product 3a in 45% yield and 99:1 er (entry 1). Replacing the counter anion (BF4−) in A with a chloride ion (pre–catalyst B) led to comparable results for this model substrate (entry 2). As an important technical note, in subsequent studies for scope explorations, we found that pre–catalyst B consistently performed better for all the substrate examinations. The N–mesityl substituent in A could be switched by a phenyl unit (pre–catalyst C) with little effect on product yield or er value (entry 3). Further optimizations with respects to bases and solvents were performed by using NHC pre–catalyst B (entries 5–12). At last, we found that by using K3PO4 as the base with CH2Cl2 as the solvent, product 3a could be isolated in 89% yield with over 99:1 er (entry 10).

Having an acceptable condition in hand, the generality of the reaction was then investigated (Fig. 2). Various substituents were placed on the para–carbon (relative to the aldehyde moiety) on the phenyl ring of 1a, in all cases the mono–ester products were obtained with excellent er values (mostly over 99:1 er, 3b to 3j). The reaction yields are good as well when the substituents are methyl (3b), methoxyl (3c), ethylthio (3d) or halogen atoms (3e–3g), giving the corresponding products with 60–94% yields. When electron–withdrawing units (e.g., CN, CF3) were used, the products (3 h, 3i) were obtained in slightly lower yields (60% and 61% yields) with excellent er values maintained. The main side products were from further esterification reaction of 3h and 3i to give the corresponding di–ester adducts. Various substituents (such as Me, OBn and halogen) could be installed on the meta–carbon (relative to the aldehyde) on the phenyl ring of 1a as well without affecting reaction yields and er values (3k–3o). Remarkably, substrates with two substituents on both the para– and meta–carbons of 1a were well tolerated (3p and 3q). When a methyl unit was placed on the ortho–position (relative to aldehyde) of 1a, drops on both reaction yield and er value were observed (3s). The low yield of 3s was mainly due to di–ester formation, and the origins for the decrease of er value may result from steric hindrance. Fluorine substituent at ortho–position led to product 3t with 80% yield and 95:5 er. Placing a methyl unit on the ortho–carbon (relative to the sulfoxide unit) of 1a led to 3r with over 99:1 er, albeit with a decreased 47% yield. Moreover, various alcohols and thiols, including secondary alcohols, could also be used as effective nucleophiles to replace methanol (3u–3x). Interestingly, when diols were used as the nucleophiles, both of hydroxyl moieties could be acylated to give the corresponding chiral di–sulfoxides with excellent yields and er values (3y, 3z). These results suggested that our strategy may be further developed to attach chiral sulfoxide to functional molecules (such as natural products and polymers) which contain multiple hydroxyl units.

aReaction conditions as stated in Table 1, entry 10. Yields are isolated yields after purification by column chromatography. Er values were determined via HPLC on chiral stationary phase. b50 °C and THF as solvent. c220 mol% 1a, DQ, K3PO4 and 100 mol% diol were used.

In synthetic applications, our approach could be readily scaled up to 1.2 grams only with little influence on product yield (e.g., 3a, 1.2 grams, 79% yield, and >99:1 er; Fig. 2). The remaining aldehyde unit in our sulfoxide product 3a could be easily converted to a diverse set of functional groups (Fig. 3a). For instance, the hydrogen of aldehyde could be deuterated61 catalyzed by achiral NHC in the presence of D2O to afford 100% deuterated 4a in 77% yield and without the loss of optical purity. Moreover, the formyl group could be cyanation62 and thioesterification catalyzed by achiral NHC with high er values (4b, 4c). Enantioenriched terminal alkyne 4d and alkene 4e were synthesized efficiently by means of Seyferth–Gilbert reaction63 and Wittig reaction64, respectively. Chiral sulfoxide 3a reacted with L–valinol65 generated oxazolines 4g was very similar to the SOX type ligands66,67 (Fig. 1e). Noteworthily, the chiral sulfoxide 4h and its analogues have been proven as a chiral ligand and catalyst in several asymmetric synthesis59. It could be easily synthesized from 3a via reductive amination reaction as well as its analogues. Furthermore, 3a underwent hydrolysis of the ester group and subsequent reductive amination of the formyl group with BnNH2 to afford an unnatural amino acid 4i bearing a chiral sulfoxide center with good yield and excellent er value. Combination of 3a with Ellman auxiliary68 accessed to a chiral disulfoxide product 4j efficiently via a concise condensation reaction with 90% yield.

aK3PO4, NBS, 4 Å MS, 30 °C, toluene; bMgSO4, 4 h, CH2Cl2, NaBH3CN; cNaBH4, Ti(OEt)4, CH2Cl2; dLiOH, THF:H2O = 2:1, 2 h, then 1 M HCl; eMgSO4, BnNH2, 4 h, then NaBH3CN; fPyrrolidine, 4 Å MS, CH2Cl2, 60 °C; gachiral NHC, AcOK, D2O:CH2Cl2 = 4:1; hachiral NHC, TsNH2, Et2NH, 4 Å MS, toluene; iachiral NHC, DQ, EtSH, K3PO4, CH2Cl2; jTMSCHN2, LDA, THF, −78 °C; kCH3PPh3Br, KHMDS, THF; lPd/C, H2, EtOH. a Synthetic transformation of 3a. b Synthetic applications.

Moreover, two practical applications of 4d and 4g were verified. Alkyne 4d could be conjugated with an anti–HIV drug (Zidovudine)69 which possessing an azido group to afford a modified Zidovudine 5a with moderate yield. As we expected, 4g could be a potential chiral ligand in asymmetric synthetic chemistry, which was used as a chiral ligand in the Pd-catalyzed enantioselective substitution reaction70 between the alkene 6 and the malonate 7, with the chiral product 8 afforded in 98:2 er. (Fig. 3b)

To understand the possible impacts of chalcogen bonding interactions, we examined two other sulfoxide substrates (1aa and 1ab) by placing the positions of aldehyde moieties which are far away from the sulfoxides sulfur center (Fig. 4a). From analysis on the X–ray structure of 1aa, the remote aldehyde unit does not show any chalcogen bondind interaction with the sulfur atom (see Supplementary Table 2). It is therefore expected that the chalcogen bond–guided conformational resolution strategy developed here shall not work for substrates such as 1aa and 1ab. This expectation was verified by our experimental observations when the use of 1aa and 1ab under our condition, It gave the corresponding products (3aa and 3ac) with nearly no enantiomeric excesses and the yields of di-esters were increased.

a Control experiments without chalcogen bonding interaction. b Quantify the strength of chalcogen bonding interaction via DFT calculations. c Addition barrier of NHC to aldehyde moieties. d Oxidized barrier of Breslow intermediates.

To probe further mechansic insights of our reactions, we investigated the chalcogen bonding strength between the sulfoxide and formyl group by using DFT calcuations (Fig. 4b). The structure provided by single crystal diffraction data of 1a was used as the initial point for geometric optimizations. The chalcogen bond energy (ChBE) was estimated to be 3.44 kcal/mol (Fig. 4b). Furthemore, in order to evalued the influence of substituents to the ChBs, substrates 1c (with OMe) and 1i (with CF3) were examined with the same DFT calculation method. The initial structures for DFT calculations were obtained from the corresponding single crystall of 1c (CCDC 2172904) and 1i (CCDC 2172911). The results showed that the chalcogen bond strength of 1c and 1i were 3.56 and 4.27 kcal/mol, respectively (see Supplementary Fig. 1 for details). Moreover, the additions of NHC catalyst to aldehyde moiety in the two conformational isomers (two set of enantiomers; four possibilities for the additions) were then evaluated (Fig. 4c). We found that the aldheyde moieties involved in chalcogen bonding interactions are conformationally locked and weakly activated. These conformationally locked aldehyde moieties react faster with the NHC catalsyt (ΔG‡ = 5.50, 3.53 kcal/mol) than the no chalcogen bonded aldehyde moieties (ΔG‡ = 6.49, 9.24 kcal/mol). Meantime, the low rotation barriers (Fig. 4b) indicates that the conformations of 1a can undergo rapid interconversions at room temperature, making it feasible to achieve a carbene–catalyzed DKR process. Further DFT studies suggest that oxidation of the Breslow intermediate (Fig. 4d) is the stereo–determine step. The activation energy difference of Ox–Ts I and Ox–Ts I′ (ΔG‡ = 14.16, 22.22 kcal/mol, respectively) is estimated as 8.06 kcal/mol, suggesting an er value over 99:1, that is consistent with our experimental observations (see Supplementary Fig. 3 for details).

In summary, we have disclosed a carbene-catalyzed DKR strategy for the synthesis of chiral sulfoxides. This method takes advantage of intramolecular chalcogen bonds installed in molecules to guide conformational isomerization and reactivity differentiation of substrates. In particular, through a chalcogen bonding–enabled reactivity differentiation, we realize a carbene–catalyzed dynamic kinetic resolution process for efficient preparation of chiral sulfoxides with excellent optical purities. The chiral sulfoxide products from our reactions may serve as platform scaffolds for straightforward transformation to useful molecules with applications in catalysis and biological studies. Chalcogen bonding interactions are naturally present or can be readily installed in various molecules. The strategy reported herein may open a new avenue in reaction control and asymmetric synthesis.

To a 100.0 mL over–dried round bottom flask equipped with a magnetic stir bar was added 1a (1.0 g, 3.87 mmol), DQ (1.58 g, 3.87 mmol), pre–NHC B (139.8 mg, 0.38 mmol) and K3PO4 (164.2 mg, 0.77 mmol). The flask was then sealed, purged and backfilled with N2 three times in glovebox before adding CH2Cl2 (60.0 mL) and CH3OH (0.19 mL, 4.65 mmol), and the reaction mixture was stirred in oil bath at 30 °C for 12 h. The mixture was concentrated under reduced pressure. The resulting crude residue was purified via column chromatography on silica gel by using petroleum ether / ethyl acetate (2:1) to afford the desired product 3a (881.8 mg, 79% yield, > 99:1 er).

The experimental method and data generated in this study are provided in the Supplementary Information file. Geometries of all DFT-optimized structures (in.xyz format) are provided as Supplementary Data file. The crystallographic data for structures of 1a, 1c, 1i, 1aa and 3a have been deposited in the Cambridge Crystallographic Data Centre under accession CCDC code 2143570, 2172904, 2172911, 2143573 and 2143579, respectively. Copies of the data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif.

Doyle, A. G. & Jacobsen, E. N. Small–molecule H–bond donors in asymmetric catalysis. Chem. Rev. 107, 5713–5743 (2007).

Article CAS PubMed Google Scholar

Banik, S. M., Levina, A., Hyde, A. M. & Jacobsen, E. N. An enantioconvergent halogenophilic nucleophilic substitution (SN2X) reaction. Science 358, 761–764 (2017).

Article ADS CAS PubMed PubMed Central Google Scholar

Wendlandt, A. E., Vangal, P. & Jacobsen, E. N. Quaternary stereocentres via an enantioconvergent catalytic SN1 reaction. Nature 556, 447–451 (2018).

Article ADS CAS PubMed PubMed Central Google Scholar

Desiraju, G. R. et al. Definition of the halogen bond. Pure. Appl. Chem. 85, 1711–1713 (2013).

Article CAS Google Scholar

Cavallo, G. et al. The halogen bond. Chem. Rev. 116, 2478–2601 (2016).

Article CAS PubMed PubMed Central Google Scholar

Chanand, Y. C. & Yeung, Y. Y. Halogen bond catalyzed bromocarbocyclization. Angew. Chem. Int. Ed. 57, 3483–3487 (2018).

Article CAS Google Scholar

Wolf, J., Huber, F. & Huber, S. M. Activation of a metal–halogen bond by halogen bonding. Angew. Chem. Int. Ed. 59, 16496–16500 (2020).

Article CAS Google Scholar

Aakeroy, C. B. et al. Definition of the chalcogen bond. Pure. Appl. Chem. 91, 1889–1892 (2019).

Article CAS Google Scholar

Vogel, L., Wonner, P. & Huber, S. M. Chalcogen bonding: an overview. Angew. Chem. Int. Ed. 58, 1880–1891 (2019).

Article CAS Google Scholar

Kolb, S., Oliver, G. A. & Werz, D. B. Chemistry evolves, terms evolve, but phenomena do not evolve: from chalcogen–chalcogen interactions to chalcogen bonding. Angew. Chem. Int. Ed. 59, 22306–22310 (2020).

Article CAS Google Scholar

Taylor, J. C. & Markham, G. D. The bifunctional active site of S–adenosylmethionine synthetase. Roles of the active site aspartates. J. Biol. Chem. 274, 32909–32914 (1999).

Article CAS PubMed Google Scholar

Horowitz, S., Dirk, L. M. A., Yesselman, J. D. & Trievel, R. C. Conservation and functional importance of carbon–oxygen hydrogen bonding in adoMet–dependent methyltransferases. J. Am. Chem. Soc. 135, 15536–15548 (2013).

Article CAS PubMed Google Scholar

Mugesh, G. et al. Glutathione peroxidase–like antioxidant activity of diaryl diselenides:  a mechanistic study. J. Am. Chem. Soc. 123, 839–850 (2001).

Article CAS PubMed Google Scholar

Werz, D. B. et al. Self–organization of chalcogen–containing cyclic alkynes and alkenes to yield columnar structures. Org. Lett. 4, 339–342 (2002).

Article CAS PubMed Google Scholar

Garrett, G. E., Gibson, G. L., Straus, R. N., Seferos, D. S. & Taylor, M. S. Chalcogen bonding in solution: interactions of benzotelluradiazoles with anionic and uncharged Lewis bases. J. Am. Chem. Soc. 137, 4126–4133 (2015).

Article CAS PubMed Google Scholar

Benz, S. et al. Anion transport with chalcogen bonds. J. Am. Chem. Soc. 138, 9093–9096 (2016).

Article CAS PubMed Google Scholar

Borissov, A., Marques, I. & Beer, P. D. Anion recognition in water by charge–neutral halogen and chalcogen bonding foldamer receptors. J. Am. Chem. Soc. 141, 4119–4129 (2019).

Article CAS PubMed Google Scholar

Cozzolino, A. F., Vargas–B, I., Mansour, S. & Mahmoudkhani, A. H. The nature of the supramolecular association of 1,2,5–chalcogenadiazoles. J. Am. Chem. Soc. 127, 3184–3190 (2005).

Article CAS PubMed Google Scholar

Gleiter, R., Haberhauer, G., Werz, D. B., Rominger, F. & Bleiholder, C. From noncovalent chalcogen–chalcogen interactions to supramolecular aggregates: experiments and calculations. Chem. Rev. 118, 2010–2041 (2018).

Article CAS PubMed Google Scholar

Chen, L., Xiang, J., Zhao, Y. & Yan, Q. Reversible self–assembly of supramolecular vesicles and nanofibers driven by chalcogen–bonding interactions. J. Am. Chem. Soc. 140, 7079–7082 (2018).

Article CAS PubMed Google Scholar

Nagao, Y., Hirata, T., Kakehi, A., Iizuka, K. & Shiro, M. Intramolecular nonbonded S…O interaction recognized in (acylimino)thiadiazoline derivatives as angiotensin II receptor antagonists and related compounds. J. Am. Chem. Soc. 120, 3104–3110 (1998).

Article CAS Google Scholar

Alamiddine, Z., Thany, S. & Le Questel, J. Y. Conformations and binding properties of thiametoxamand clothianidin neonicotinoid insecticides to nicotinic acetylcholine receptors: the contribution of σ–hole interactions. Chem. Phys. Chem. 19, 3069–3083 (2018).

Article CAS PubMed Google Scholar

Thomas, S. P., Jayatilaka, D. & Guru, R. T. N. S…O chalcogen bonding in sulfa drugs: insights from multipole charge density and X–ray wavefunction of acetazolamide. Phys. Chem. Chem. Phys. 17, 25411–25420 (2015).

Article CAS PubMed Google Scholar

Burling, F. T. & Goldstein, B. M. Computational studies of nonbonded sulfur–oxygen and selenium–oxygen interactions in the thiazole and selenazole nucleosides. J. Am. Chem. Soc. 114, 2313–2320 (1992).

Article CAS Google Scholar

Young, C. M. et al. The importance of 1,5–oxygen···chalcogen interactions in enantioselective isochalcogenourea catalysis. Angew. Chem. Int. Ed. 59, 3705–3710 (2020).

Article CAS Google Scholar

Weiss, R., Aubert, E., Pale, P. & Mamane, V. Chalcogen–bonding catalysis with telluronium cations. Angew. Chem. Int. Ed. 60, 19281–19286 (2021).

Article CAS Google Scholar

He, X., Wang, X., Tse, Y. L., Ke, Z. & Yeung, Y. Y. Bis–selenonium cations as bidentate chalcogen bond donors in catalysis. ACS Catal. 11, 12632–12642 (2021).

Article CAS Google Scholar

Benz, S., Andarias, J. L. & Matile, S. Catalysis with chalcogen bonds. Angew. Chem. Int. Ed. 56, 812–815 (2017).

Article CAS Google Scholar

Benz, S. & Matile, S. Catalysis with pnictogen, chalcogen, and halogen bonds. Angew. Chem. Int. Ed. 57, 5408–5412 (2018).

Article CAS Google Scholar

Wonner, P., Vogel, L., Werz, D. B. & Huber, S. M. Carbon–halogen bond activation by selenium–based chalcogen bonding. Angew. Chem. Int. Ed. 56, 12009–12012 (2017).

Article CAS Google Scholar

Wonner, P., Dreger, A. & Huber, S. M. Chalcogen bonding catalysis of a nitro–Michael reaction. Angew. Chem. Int. Ed. 58, 16923–16927 (2019).

Article CAS Google Scholar

Wang, W. et al. Chalcogen−chalcogen bonding catalysis enables assembly of discrete molecules. J. Am. Chem. Soc. 141, 9175–9179 (2019).

Article CAS PubMed Google Scholar

Wang, W. et al. Dual chalcogen−chalcogen bonding catalysis. J. Am. Chem. Soc. 142, 3117–3124 (2020).

Article CAS PubMed Google Scholar

Kong, X., Zhou, P. P. & Wang, Y. Chalcogen···π bonding catalysis. Angew. Chem. Int. Ed. 60, 9395–9400 (2021).

Article CAS Google Scholar

Iwaoka, M. & Tomoda, S. Nature of the intramolecular Se…N nonbonded interaction of 2–selenobenzylamine derivatives. An experimental evaluation by 1H, 77Se, and 15N NMR Spectroscopy. J. Am. Chem. Soc. 118, 8077–8084 (1996).

Article CAS Google Scholar

Wu, X. X. et al. Catalyst control over sixfold stereogenicity. Nat. Catal. 4, 457–462 (2021).

Article CAS Google Scholar

Fujita, K., Iwaoka, M. & Tomoda, S. Synthesis of diaryl diselenides having chiral pyrrolidine rings with C2 symmetry. Their application to the asymmetric methoxyselenenylation of trans-β-methylstyrenes. Chem. Lett. 23, 923–926 (1994).

Article Google Scholar

Worth, T. Asymmetric reaction of arylalkenes with diselenides. Angew. Chem. Int. Ed. Engl. 34, 1726–1728 (1995).

Article Google Scholar

Arokianathar, J. N., Frost, A. B. & Smith, A. D. Isothiourea-catalyzed enantioselective addition of 4-nitrophenyl esters to iminium ions. ACS Catal. 8, 1153–1160 (2018).

Article CAS Google Scholar

Fukata, Y., Asano, K. & Matsubara, S. Facile net cycloaddition approach to optically active 1,5-benzothiazepines. J. Am. Chem. Soc. 137, 5320–5323 (2015).

Article CAS PubMed Google Scholar

Belmessieri, D. et al. Organocatalytic functionalization of carboxylic acids: isothiourea-catalyzed asymmetric intra- and intermolecular Michael addition-lactonizations. J. Am. Chem. Soc. 133, 2714–2720 (2011).

Article CAS PubMed Google Scholar

Leverett, C. A., Purohit, V. C. & Romo, D. Enantioselective, organocatalyzed, intramolecular aldol lactonizations with keto acids leading to bi- and tricyclic β-lactones and topology-morphing transformations. Angew. Chem. Int. Ed. 49, 9479–9483 (2010).

Article CAS Google Scholar

Robinson, E. R., Fallan, C. & Smith, A. D. Anhydrides as α, β-unsaturated acyl ammonium precursors: isothiourea-promoted catalytic asymmetric annulation processes. Chem. Sci. 4, 2193–2200 (2013).

Article CAS Google Scholar

Yang, X. & Birman, V. B. Kinetic resolution of α-substituted alkanoic acids promoted by homobenzotetramisole. Chem. Eur. J. 17, 11296–1130 (2011).

Article CAS PubMed Google Scholar

Mukherjee, A. J., Zade, S. S., Singh, H. B. & Sunoj, R. B. Organoselenium chemistry: role of intramolecular interactions. Chem. Rev. 110, 4357–4416 (2010).

Article CAS PubMed Google Scholar

Srivastava, K., Shah, P., Singh, H. B. & Butcher, R. J. Isolation and structural characterization of some aryltellurium halides and their hydrolyzed products stabilized by an intramolecular Te-N interaction. Organometallics 30, 534–546 (2011).

Article CAS Google Scholar

Srivastava, K., Chakraborty, T., Singh, H. B. & Butcher, R. J. Intramolecularly coordinated azobenzene selenium derivatives: Effect of strength of the Se-N intramolecular interaction on luminescence. Dalton Trans. 40, 4489–4496 (2011).

Article CAS PubMed Google Scholar

Tiecco, M. et al. Preparation of a new chiral non-racemic sulfur-containing diselenide and applications in asymmetric Synthesis. Chem. Eur. J. 8, 1118–1124 (2002).

3.0.CO;2-2" data-track-action="article reference" href="https://doi.org/10.1002%2F1521-3765%2820020301%298%3A5%3C1118%3A%3AAID-CHEM1118%3E3.0.CO%3B2-2" aria-label="Article reference 48" data-doi="10.1002/1521-3765(20020301)8:53.0.CO;2-2">Article CAS PubMed Google Scholar

Echeverria, J. Cooperative effects between hydrogen bonds and C=O···S interactions in the crystal structures of sulfoxides. Cryst. Growth Des. 21, 2481–2487 (2021).

Article CAS Google Scholar

Enders, D., Niemeier, O. & Henseler, A. Organocatalysis by N–heterocyclic carbenes. Chem. Rev. 107, 5606–5655 (2007).

Article CAS PubMed Google Scholar

Hopkinson, M. N., Richter, C. & Glorius, F. An overview of N–heterocyclic carbenes. Nature 510, 485–496 (2014).

Article ADS CAS PubMed Google Scholar

Mahatthananchai, J. & Bode, J. W. On the mechanism of N–heterocyclic carbene–catalyzed reactions involving acyl azoliums. Acc. Chem. Res. 47, 696–707 (2014).

Article CAS PubMed Google Scholar

Flanigan, D. M., White, N. A. & Rovis, T. Organocatalytic reactions enabled by N‑heterocyclic carbenes. Chem. Rev. 115, 9307–9387 (2015).

Article CAS PubMed PubMed Central Google Scholar

Murauski, K. J. R., Jaworski, A. A. & Scheidt, K. A. A continuing challenge: N–heterocyclic carbene–catalyzed syntheses of γ–Butyrolactones. Chem. Soc. Rev. 47, 1773–1782 (2018).

Article CAS PubMed Google Scholar

Olbe, L., Carlsson, E. & Lindberg, P. A proton–pump inhibitor expedition: the case histories of omeprazole and esomeprazole. Nat. Rev. Drug. Discov. 2, 132–139 (2003).

Article CAS PubMed Google Scholar

Garnock–Jones, K. P., Dhillon, S. & Scott, L. J. Armodafinil. Cns. Drugs 23, 793–803 (2009).

Article PubMed Google Scholar

Caboni, P., Sammelson, R. E. & Casida, J. E. Phenylpyrazole insecticide photochemistry, metabolism, and GABAergic action:  ethiprole compared with fipronil. J. Agric. Food Chem. 51, 7055–7061 (2003).

Article CAS PubMed Google Scholar

Sipos, G., Drinkel, E. E. & Dorta, R. The emergence of sulfoxides as efficient ligands in transition metal catalysis. Chem. Soc. Rev. 44, 3834–3860 (2015).

Article CAS PubMed Google Scholar

Otocka, S., Kwiatkowska, M., Madalinska, L. & Kiełbasinski, P. Chiral organosulfur ligands/catalysts with a stereogenic sulfur atom: applications in asymmetric synthesis. Chem. Rev. 117, 4147–4181 (2017).

Article CAS PubMed Google Scholar

Sarkar, S. D. & Studer, A. NHC–catalyzed Michael addition to α, β–unsaturated aldehydes by redox activation. Angew. Chem. Int. Ed. 49, 9266–9269 (2010).

Article CAS Google Scholar

Geng, H. et al. Practical synthesis of C1 deuterated aldehydes enabled by NHC catalysis. Nat. Catal. 2, 1071–1077 (2019).

Article CAS PubMed PubMed Central Google Scholar

Lv, Y. et al. Catalytic atroposelective synthesis of axially chiral benzonitriles via chirality control during bond dissociation and CN group formation. Nat. Commun. 13, 36 (2022).

Article ADS CAS PubMed PubMed Central Google Scholar

Seyferth, D., Hilbert, P. & Marmor, R. S. Novel diazo alkanes and the first carbene containing the dimethyl phosphite group. J. Am. Chem. Soc. 89, 4811–4812 (1967).

Article CAS Google Scholar

Wittig, G. & Schöllkopf, U. Über triphenyl–phosphin–methylene als olefinbildende reagenzien. Chem. Ber. 87, 1318–1330 (1954).

Article Google Scholar

Schwekendiek, K. & Glorius, F. Efficient oxidative synthesis of 2–oxazolines. Synthesis 18, 2996–3002 (2006).

Google Scholar

Liu, W., Ali, S. Z., Ammann, S. E. & White, M. C. Asymmetric allylic C−H alkylation via palladium (II)/cis–ArSOX catalysis. J. Am. Chem. Soc. 140, 10658–10662 (2018).

Article CAS PubMed PubMed Central Google Scholar

Ammann, S. E., Liu, W. & White, M. C. Enantioselective allylic C–H oxidation of terminal olefins to isochromans by palladium (II)/chiral sulfoxide catalysis. Angew. Chem. Int. Ed. 55, 9571–9575 (2016).

Article CAS Google Scholar

Robak, M., Herbage, M. A. & Ellman, J. A. Synthesis and applications of tert–butanesulfinamide. Chem. Rev. 110, 3600–3740 (2010).

Article CAS PubMed Google Scholar

McLeod, G. X. & Hammer, S. M. Zidovudine: five years later. Ann. Intern. Med. 117, 487–501 (1992).

Article CAS PubMed Google Scholar

Du, L., Cao, P. & Liao, J. Hydrogen–bond–promoted palladium catalysis: Allylic alkylation of indoles with unsymmetrical 1,3–disubstituted allyl acetates using chiral bis(sulfoxide) phosphine ligands. Angew. Chem. 125, 4301–4305 (2013).

Article ADS Google Scholar

Download references

We acknowledge funding supports from National Natural Science Foundation of China (21732002, 22061007, P. C. Z; 22071036, Y. R. C); Frontiers Science Center for Asymmetric Synthesis and Medicinal Molecules, Department of Education, Guizhou Province [Qianjiaohe KY number (2020)004, Y. R. C]; The 10 Talent Plan (Shicengci) of Guizhou Province ([2016]5649, Y. R. C); Qiankehejichu-ZK[2022]zhongdian024, P. C. Z); Science and Technology Department of Guizhou Province ([2018]2802, [2019]1020, Y. R. C); Program of Introducing Talents of Discipline to Universities of China (111 Program, D20023, Y. R. C) at Guizhou University. Singapore National Research Foundation under its NRF Investigatorship (NRF–NRFI2016–06, Y. R. C) and Competitive Research Program (NRF–CRP22–2019–0002, Y. R. C); Ministry of Education, Singapore, under its MOE AcRF Tier 1 Award (RG7/20, RG5/19, Y. R. C), MOE AcRF Tier 2 (MOE2019–T2–2–117, Y. R. C), and MOE AcRF Tier 3 Award (MOE2018–T3–1–003, Y. R. C), Nanyang Technological University.

State Key Laboratory Breeding Base of Green Pesticide and Agricultural Bioengineering, Key Laboratory of Green Pesticide and Agricultural Bioengineering, Ministry of Education, Guizhou University, Guiyang, 550025, China

Jianjian Liu, Mali Zhou, Rui Deng, Pengcheng Zheng & Yonggui Robin Chi

Division of Chemistry & Biological Chemistry, School of Physical & Mathematical Sciences, Nanyang Technological University, Singapore, 637371, Singapore

Yonggui Robin Chi

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

Y.R.C and P.C.Z conceptualized and directed this research; J.J.L designed and performed main methodology development, scope evaluation and synthetic application; M.L.Z and R.D synthesized the substrates; P.C.Z conducted the DFT calculations. All authors contributed to discussions and manuscript preparation.

Correspondence to Pengcheng Zheng or Yonggui Robin Chi.

The authors declare no competing interests.

Nature Communications thanks the anonymous reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

Liu, J., Zhou, M., Deng, R. et al. Chalcogen bond-guided conformational isomerization enables catalytic dynamic kinetic resolution of sulfoxides. Nat Commun 13, 4793 (2022). https://doi.org/10.1038/s41467-022-32428-4

Download citation

Received: 02 March 2022

Accepted: 01 August 2022

Published: 15 August 2022

DOI: https://doi.org/10.1038/s41467-022-32428-4

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

Nature Communications (2023)

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.