Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2022 Jun 1;144(23):10156–10161. doi: 10.1021/jacs.2c04261

Organocatalytic Asymmetric Synthesis of Si-Stereogenic Silyl Ethers

Hui Zhou , Jung Tae Han , Nils Nöthling , Monika M Lindner , Judith Jenniches , Clemens Kühn , Nobuya Tsuji §, Li Zhang , Benjamin List †,§,*
PMCID: PMC9490845  PMID: 35649270

Abstract

graphic file with name ja2c04261_0009.jpg

Functionalized enantiopure organosilanes are important building blocks with applications in various fields of chemistry; nevertheless, asymmetric synthetic methods for their preparation are rare. Here we report the first organocatalytic enantioselective synthesis of tertiary silyl ethers possessing “central chirality” on silicon. The reaction proceeds via a desymmetrizing carbon–carbon bond forming silicon–hydrogen exchange reaction of symmetrical bis(methallyl)silanes with phenols using newly developed imidodiphosphorimidate (IDPi) catalysts. A variety of enantiopure silyl ethers was obtained in high yields with good chemo- and enantioselectivities and could be readily derivatized to several useful chiral silicon compounds, leveraging the olefin functionality and the leaving group nature of the phenoxy substituent.


Chiral molecules bearing a carbon stereogenic center are ubiquitous in nature and have been the focus of fundamental and applied studies during the past decades.1 In contrast, the asymmetric synthesis of their heavier congeners bearing a stereogenic silicon atom has been far less investigated. As enantiopure organosilanes are currently gaining substantial importance in material science, in polymer synthesis, as chiral ligands, and in scent and medicinal chemistry, an expansion of the synthetic toolbox to enable their preparation is highly desirable.28 A literature survey reveals that Si-stereogenic silanes have been created either via desymmetrization and diastereoselective synthesis from achiral silicon compounds or via the kinetic resolution of racemic silanes by means of transition metal and enzyme catalysis.913 Indeed, the catalytic construction of Si-stereogenic centers appears to be a topic of high current relevance in asymmetric catalysis.1424 However, to the best of our knowledge, organocatalytic asymmetric approaches to enantiopure silanes constitute an unmet challenge.graphic file with name ja2c04261_0001.jpg Inspired by previous work on asymmetric catalysis of the silicon–hydrogen exchange reaction,25,26 we considered a catalytic desymmetrizing silyl ether formation from bis(methallyl)silanes 1 with alcohols 2, in which olefin protonation occurs from an acid catalyst HX* (3) to form a β-silyl-stabilized cationic intermediate, followed by nucleophilic attack of the alcohol to generate the chiral silyl ether 5 (eq 1).

We report here that instead of the intermolecular addition of the alcohol, an intramolecular cation−π cyclization with the second olefin occurs, ultimately leading to a silylium ion equivalent, which reacts with the alcohol to form an enantioenriched silyl ether 4, featuring a new C–C σ-bond (eq 2). We have discovered a highly enantioselective, desymmetrizing, carbon–carbon bond forming reaction of symmetrical bis(methallyl)silanes with phenols. The perfectly atom-economic reaction is catalyzed by newly developed imidodiphosphorimidates (IDPi) and delivers several structurally distinct silyl ethers. We also describe the utilization of the obtained products in the synthesis of various Si-stereogenic silanes and suggest a mechanism of this unusual transformation.

As the starting point, bis(methallyl)silane 1a and 2,6-dimethyl-phenol 2a were reacted with IDPi catalyst 3a, featuring a 1-naphthyl substituent at the 3,3′-position of the BINOL backbone and a trifluoromethyl sulfonyl group in the inner core. Product 4a was obtained as the main product with only small amounts of ether 5a (Table 1). Notably, we obtained the new chiral organosilane 4a in 73% yield and a promising enantioselectivity of 64:36 e.r. (Table 1, entry 1). A subsequent investigation of the reaction conditions revealed toluene as the optimal solvent, affording the desired product in 86% yield and 86:14 e.r. (entries 2–4). Lowering the temperature to −20 °C suppressed the formation of side product 5a and led to an increased yield without significantly affecting the enantioselectivity (entry 5). Spirocyclic-fluorenyl-substituted IDPi catalysts, which have been preferred motifs in our previous Si-ACDC studies, were subsequently investigated (entries 6–9).2736 We eventually identified IDPi 3e as the optimal catalyst, which enabled quantitative formation of product 4a with 97:3 e.r. (entry 9).

Table 1. Reaction Developmenta.

graphic file with name ja2c04261_0006.jpg

entry catalyst solvent T (°C) yield (%)b e.r.c
1 3a diethyl ether 25 73 64:36
2 3a dichloromethane 25 76 86:14
3 3a cyclohexane 25 90 79:21
4 3a toluene 25 86 86:14
5 3a toluene –20 94 86.5:13.5
6 3b toluene –20 >95 85:15
7 3c toluene –20 >95 89.5:10.5
8 3d toluene –20 >95 95:5
9 3e toluene –20 >95 97:3
a

Performed with 2,6-dimethylphenol 2a (0.025 mmol), 1a (1.5 equiv), and IDPi catalysts 3a3e (2.5 mol %) in solvent (0.25 mL, 0.1 M).

b

Yields determined by 1H NMR spectroscopy using 1,3,5-trimethoxybenzene as internal standard.

c

Enantiomeric ratios (e.r.) determined by HPLC.

Having identified these optimized conditions, we investigated other substrates for this new catalytic asymmetric approach to Si-stereogenic organosilanes. As illustrated in Table 2, benzyl-substituted chiral silane 4a was isolated in 92% yield with 97:3 e.r. on a 0.2 mmol scale. Related silanes 4b4d were prepared in similarly good yields and enantioselectivities (87–96% yield and 96:4 to 98:2 e.r., respectively) from the corresponding benzyl silanes with electron-donating or -withdrawing groups at the para-position of the phenyl ring. In addition, substituents with diverse electronic properties in meta- and ortho-positions of the phenyl group were well tolerated, affording chiral organosilanes 4e4i with comparable outcome. 3,4-Dimethyl-substituted product 4j was obtained in 93% yield and an e.r. of 93.5:6.5. The use of substrate 1k bearing a bulkier 2-naphthylmethyl group on the silicon atom delivered 4k with good yield and 95:5 e.r. While the reactions of benzyl-substituted silanes (1a1k) occurred with good chemoselectivity, aryl-substituted substrates (1l1n) initially led to the formation of a side product and required further optimization of the reaction conditions (see the Supporting Information, Table S3). Ultimately, this gave products 4l4n in 78–84% yields with 95:5 e.r. The silicon-stereogenic silane 4o, bearing a thiophenyl moiety, could also be obtained in satisfactory yield and with high enantioselectivity. Product 4p was generated from silane 1p, bearing three methallyl groups, with 76:24 e.r., and no further substitution was detected. Submitting silane 1q to the reaction conditions revealed the key role of the methyl substituent on the silicon atom; indeed, its replacement with an ethyl group delivered product 4q in quantitative yield but with a significantly decreased enantioselectivity (70.5:29.5 e.r.).

Table 2. Substrate Scopea.

graphic file with name ja2c04261_0007.jpg

graphic file with name ja2c04261_0008.jpg

a

Performed with 2,6-dimethylphenol 2a (0.2 mmol), 1 (1.5 equiv), and IDPi catalyst 3a or 3e (2.5 mol %) in toluene (2.0 mL, 0.1 M) at −20 °C for 24 h. Isolated yields with enantiomeric ratio (e.r.) determined by HPLC analysis.

b

With IDPi 3e.

c

With IDPi 3a.

Despite extensive attempts, we were unable to obtain suitable single crystals of our new organosilicon compounds for the determination of their absolute configuration by single-crystal X-ray diffraction. Therefore, the “crystalline sponge method” was applied to elucidate the absolute structure by X-ray analysis (CS-XRD).3740

In this method, single crystals of a flexible MOF, the crystalline sponge (CS), are used as a scaffold to arrange the analyte molecules into the pores and apply X-ray crystallographic analysis to determine the structure. With this method, we could unequivocally determine the absolute structure of the analyte molecules. The space group symmetry of the original CS changed from centrosymmetric (C2/c) to non-centrosymmetric (C2), to accommodate enantiopure molecules.

As depicted in Figure 1a, the absolute configuration of product 4b was assigned to be R by CS-XRD. Several independent additional experiments using the same and the opposite enantiomer confirmed the correctness of this assignment (see the Supporting Information, Figures S13–S30 and Tables S4–S6).

Figure 1.

Figure 1

Absolute configuration determination of (R)-4b. (a) The crystalline sponge method was employed. (b) Calculated (blue curve) and experimental (red curve) CD spectra of 4b.

The absolute configuration of products 4b and 4d was furthermore confirmed to be R by computational and experimental CD spectroscopy (Figure 1b and Figure S33 in the Supporting Information).41

To illustrate the practical utility of our methodology and the synthetic value of the newly obtained enantiopure organosilicon products, a preparative-scale reaction was performed and readily delivered 1.06 g of product 4a in 96% yield and 95:5 e.r. (Figure 2). Hydroboration/oxidation of olefin 4a gave compound 6, with moderate diastereoselectivity and without erosion of enantiopurity.42 Upon cyclopropanation or epoxidation, silane products 7 and 8 could be obtained in 40% and 55% yield, respectively, with identical e.r.43,44 Hydrogenation of 4a led to product 9 in 88% yield. Moreover, a sequential strategy was employed for the conversion of the Si–O bond into a Si–C bond. Accordingly, in situ treatment of product 9 with diisobutylaluminum hydride gave hydrosilane 10.45 A subsequent Pt-catalyzed hydrosilylation with 1-octene provided quaternary silane 11.18 While the absolute stereochemistry of this product remains to be confirmed, it has previously been shown that the reduction of alkoxysilanes with diisobutylaluminum hydride and also the hydrosilylation of 1-octene, respectively, proceed via retention.46,47 The direct construction of a Si–C bond by treatment of silyl ether 4a with n-BuLi occurred with substantial loss of enantioselectivity and gave quaternary silane 12, mostly with retention of configuration at the silicon stereocenter.48 A Pt-catalyzed intramolecular hydrosilylation of unsaturated hydrosilane 13, which is readily obtained via reduction of 4a, gave product 14 with moderate d.r. and retention of configuration.46 The Si–H bond in silane 13 could be converted into a Si–OH group via dehydrogenative coupling with water in the presence of Pd/C.49 This transformation has been shown to proceed with inversion of configuration, furnishing chiral silanol 15.50 Finally, boronate 16 was prepared in 90% yield as a 1:1 mixture of diastereomers via hydroboration.51

Figure 2.

Figure 2

Derivatizations. Isolated yields with e.r. determined by HPLC and d.r. measured by 1H NMR spectroscopy or HPLC. (a) BH3SMe2 (1.0 equiv), 0 °C, THF, 1 h, then H2O2 (7.8 equiv), rt, EtOH, 3 h. (b) Et2Zn (2.0 equiv), CH2I2 (4.0 equiv), 0 °C–rt, DCE, 18 h. (c) m-CPBA (2.0 equiv), 0 °C–rt, DCM, 7 h. (d) Pd/C (0.1 equiv), H2 (balloon), rt, MeOH, 12 h. (e) Pd/C (0.1 equiv), H2 (balloon), rt, MeOH, 12 h, then DIBAL-H (2.0 equiv), 0 °C–rt, hexanes, 18 h. (f) Pt(dvds) (0.05 equiv), 1-octene (2.0 equiv), 50 °C, hexanes, 48 h. (g) n-BuLi (3.0 equiv), 0–35 °C, Et2O, 48 h. (h) DIBAL-H (2.0 equiv), 0 °C–rt, hexanes, 18 h. (i) Pt(dvds) (0.05 equiv), 50 °C, hexanes, 48 h. (j) Pd/C (0.1 equiv), H2O (3.0 equiv), 0 °C, ethyl acetate, 24 h. (k) n-BuLi (0.1 equiv), HBPin (3.0 equiv), 130 °C, toluene, 18 h.

To elucidate the reaction mechanism, we conducted several additional experiments. Using phenyl-substituted silane 1l as starting material, we could isolate side product 5b, which enabled us to rule out its possible role as an intermediate, undergoing an intermolecular hydroallylation. As expected, mixing compounds 5b and 1l with 1.1 equiv of phenol 2b and IDPi 3a as catalyst in toluene at room temperature for 24 h did not lead to any detectable amounts of product 4l (Figure 3, eq 1). Further, when the reaction between substrates 1a and 2a was conducted for only 12 h at a reduced catalyst loading (1 mol %), cyclic Si-stereogenic silane 17a was isolated in 20% yield. Reaction progress kinetic studies suggested silane 17a to be a (“parasitic”) intermediate of the reaction (see the Supporting Information, Figure S5). Interestingly, the e.r. of this six-membered silane product was determined to be 53:47 (eq 2). That cyclization is indeed not the enantio-determining step was then confirmed when we reacted racemic silane 17a with phenol 2a, which cleanly furnished product 4a in 75% yield and 96:4 e.r. (Figure 3, eq 3). As such, the enantio-determining step is supposed to be the Si–O bond formation, and steric effects on enantiocontrol in the corresponding transition states were elucidated by DFT studies (see the Supporting Information, Figures S31 and S32 and Table S7). Furthermore, in light of the formation of product 4p from the corresponding tris-methallyl silane 1p, with significant enantioselectivity, olefin protonation as the enantio-determining step is also unlikely.

Figure 3.

Figure 3

Control experiments. Reaction of eq 1 was performed with 5b (0.1 mmol), silane 1l (1.5 equiv), phenol 2b (1.1 equiv), and IDPi 3a (2.5 mol %) in toluene (1.0 mL, 0.1 M) at rt for 24 h. Reaction of eq 2 was performed with 2a (0.2 mmol), silane 1a (1.5 equiv), and IDPi 3e (1.0 mol %) in toluene (2.0 mL, 0.1 M) at −20 °C for 12 h and terminated by the addition of Et3N. Reaction of eq 3 was performed with rac-17a (0.2 mmol), 2a (1.1 equiv), and IDPi 3e (2.5 mol %) in toluene (2.0 mL, 0.1 M) at −20 °C for 24 h.

Based on these results, a plausible reaction mechanism can be proposed (Figure 4). Accordingly, the catalytic cycle commences with the protonation of symmetrical silane 1 by IDPi 3 to provide ion pair I, the carbocation of which is stabilized by silicon hyperconjugation. Subsequent cation−π cyclization takes place to afford the ion pair II. Deprotonation of its cyclic cation gives compound 17, an isolable intermediate that can reversibly be protonated to regenerate ion pair II. Alternatively, Si–C bond cleavage would lead to silylium-based ion pair III. Finally, reaction of this intermediate with phenol 2a furnishes product 4 and regenerates catalyst 3.

Figure 4.

Figure 4

Proposed mechanism.

In conclusion, we have realized an organocatalytic asymmetric synthesis of Si-stereogenic silyl ethers that proceeds via a C–C bond forming desymmetrization and is enabled by our IDPi catalysts. Various non-natural, enantioenriched silane products could be generated and were utilized in the synthesis of valuable silane derivatives with potential application in material and medicinal chemistry. Our approach features scalability, broad substrate scope, operational simplicity, and mechanistic novelty. Particularly, we observed and characterized an unprecedented six-membered cyclic chiral silane, the protonated form of which may act as an intermediate in the catalytic cycle. Our newly developed strategy provides a practical and efficient access to Si-stereogenic compounds, which may find utilization in the synthesis of silicon-containing materials, pharmaceuticals, and chiral ligands for transition-metal-catalyzed reactions.

Acknowledgments

Generous support from the Deutsche Forschungsgemeinschaft (Leibniz Award to B.L. and Germany’s Excellence Strategy-EXC 2033-390677874-RESOLV) and the European Research Council (European Union’s Horizon 2020 research and innovation program “C–H Acids for Organic Synthesis, CHAOS” Advanced Grant Agreement No. 694228) is gratefully acknowledged. This work was also financially supported by the Institute for Chemical Reaction Design and Discovery (ICReDD), which was established by the World Premier International Research Initiative (WPI), MEXT, Japan, and by JSPS KAKENHI Grants 21H01925 and 20K22515. The authors thank Dr. Roberta Properzi and Prof. Dr. Christian Lehmann for helpful discussions and Dr. Markus Leutzsch for NMR studies. We thank DESY (Hamburg, Germany), a member of the Helmholtz Association HGF, for the provision of experimental facilities; parts of this research were carried out at PETRA III, and we would like to thank Sofiane Saouane for excellent assistance in using the P11-High-throughput Macromolecular Crystallography Beamline. We also appreciate the support by the technicians of our group and thank the members of our MS and chromatography groups for their excellent service.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.2c04261.

  • Experimental details and analytical data for all new compounds (PDF)

Accession Codes

CCDC 2115183 and 21664022166408 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

Open access funded by Max Planck Society.

The authors declare the following competing financial interest(s): We have a patent on IDPi catalysts and their use in asymmetric catalysis.

Supplementary Material

ja2c04261_si_001.pdf (19.4MB, pdf)

References

  1. Carreira E. M.; Yamamoto H.. Comprehensive Chirality; Elsevier: Amsterdam, 2012. [Google Scholar]
  2. Tacke R.; Linoh H.. Bioorganosilicon Chemistry. Organic Silicon Compounds; J. Wiley and Sons: New York, 1989. [Google Scholar]
  3. Magnus P.Silicon in Organic, Organometallic, and Polymer Chemistry; J. Wiley and Sons: New York, 2000. [Google Scholar]
  4. Franz A. K.; Wilson S. O. Organosilicon Molecules with Medicinal Applications. J. Med. Chem. 2013, 56, 388–405. 10.1021/jm3010114. [DOI] [PubMed] [Google Scholar]
  5. Shintani R.; Takano R.; Nozaki K. Rhodium-Catalyzed Asymmetric Synthesis of Silicon-Stereogenic Silicon-Bridged Arylpyridinones. Chem. Sci. 2016, 7, 1205–1211. 10.1039/C5SC03767K. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Shintani R.; Misawa N.; Takano R.; Nozaki K. Rhodium-Catalyzed Synthesis and Optical Properties of Silicon-Bridged Arylpyridines. Chem.—Eur. J. 2017, 23, 2660–2665. 10.1002/chem.201605000. [DOI] [PubMed] [Google Scholar]
  7. Bai X. F.; Zou J. F.; Chen M. Y.; Xu Z.; Li L.; Cui Y. M.; Zheng Z. J.; Xu L. W. Lewis-Base-Mediated Diastereoselective Silylations of Alcohols: Synthesis of Silicon-Stereogenic Dialkoxysilanes Controlled by Chiral Aryl BINMOLs. Chem.—Asian J. 2017, 12, 1730–1735. 10.1002/asia.201700640. [DOI] [PubMed] [Google Scholar]
  8. Ramesh R.; Reddy D. S. Quest for Novel Chemical Entities through Incorporation of Silicon in Drug Scaffolds. J. Med. Chem. 2018, 61, 3779–3798. 10.1021/acs.jmedchem.7b00718. [DOI] [PubMed] [Google Scholar]
  9. Oestreich M. Silicon-Stereogenic Silanes in Asymmetric Catalysis. Synlett 2007, 2007, 1629–1643. 10.1055/s-2007-980385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. a Xu L. W.; Li L.; Lai G. Q.; Jiang J. X. The Recent Synthesis and Application of Silicon-Stereogenic Silanes: A Renewed and Significant Challenge in Asymmetric Synthesis. Chem. Soc. Rev. 2011, 40, 1777–1790. 10.1039/C0CS00037J. [DOI] [PubMed] [Google Scholar]
  11. Xu L. W. Desymmetrization Catalyzed by Transition Metal Complex: Enantioselective Construction of Silicon-Stereogenic Silanes. Angew. Chem., Int. Ed. 2012, 51, 12932–12934. 10.1002/anie.201207932. [DOI] [PubMed] [Google Scholar]
  12. Igawa K.; Tomooka K. Chiral Silicon Molecules. Organosilicon Chemistry: Novel Approaches and Reactions 2019, 495–532. 10.1002/9783527814787.ch14. [DOI] [Google Scholar]
  13. Ye F.; Xu Z.; Xu L. W. The Discovery of Multifunctional Chiral P Ligands for the Catalytic Construction of Quaternary Carbon/Silicon and Multiple Stereogenic Centers. Acc. Chem. Res. 2021, 54, 452–470. 10.1021/acs.accounts.0c00740. [DOI] [PubMed] [Google Scholar]
  14. Oestreich M.; Rendler S. True Chirality Transfer from Silicon to Carbon: Asymmetric Amplification in a Reagent-Controlled Palladium-Catalyzed Hydrosilylation. Angew. Chem., Int. Ed. 2005, 44, 1661–1664. 10.1002/anie.200462355. [DOI] [PubMed] [Google Scholar]
  15. Nakazaki A.; Nakai T.; Tomooka K. Asymmetric Retro-[1,4] Brook Rearrangement and Its Stereochemical Course at Silicon. Angew. Chem., Int. Ed. 2006, 45, 2235–2238. 10.1002/anie.200503734. [DOI] [PubMed] [Google Scholar]
  16. Shintani R.; Moriya K.; Hayashi T. Palladium-Catalyzed Enantioselective Desymmetrization of Silacyclobutanes: Construction of Silacycles Possessing a Tetraorganosilicon Stereocenter. J. Am. Chem. Soc. 2011, 133, 16440–16443. 10.1021/ja208621x. [DOI] [PubMed] [Google Scholar]
  17. Bauer J. O.; Strohmann C. Stereoselective Synthesis of Silicon-Stereogenic Aminomethoxysilanes: Easy Access to Highly Enantiomerically Enriched Siloxanes. Angew. Chem., Int. Ed. 2014, 53, 720–724. 10.1002/anie.201307826. [DOI] [PubMed] [Google Scholar]
  18. Zhan G.; Teng H. L.; Luo Y.; Lou S.-J.; Nishiura M.; Hou Z. M. Enantioselective Construction of Silicon-Stereogenic Silanes by Scandium-Catalyzed Intermolecular Alkene Hydrosilylation. Angew. Chem., Int. Ed. 2018, 57, 12342–12346. 10.1002/anie.201807493. [DOI] [PubMed] [Google Scholar]
  19. Zhang Q.-W.; An K.; Liu L.-C.; Zhang Q.; Guo H.; He W. Construction of Chiral Tetraorganosilicons by Tandem Desymmetrization of Silacyclobutanes/Intermolecular Dehydrogenative Silylation. Angew. Chem., Int. Ed. 2017, 56, 1125–1129. 10.1002/anie.201609022. [DOI] [PubMed] [Google Scholar]
  20. Wen H. A.; Wan X. L.; Huang Z. Asymmetric Synthesis of Silicon-Stereogenic Vinylhydrosilanes by Cobalt-Catalyzed Regio- and Enantioselective Alkyne Hydrosilylation with Dihydrosilanes. Angew. Chem., Int. Ed. 2018, 57, 6319–6323. 10.1002/anie.201802806. [DOI] [PubMed] [Google Scholar]
  21. Chen H.; Chen Y.; Tang X.; Liu S.; Wang R.; Hu T.; Gao L.; Song Z. Rhodium-Catalyzed Reaction of Silacyclobutanes with Unactivated Alkynes to Afford Silacyclohexenes. Angew. Chem., Int. Ed. 2019, 58, 4695–4699. 10.1002/anie.201814143. [DOI] [PubMed] [Google Scholar]
  22. Jagannathan J. R.; Fettinger J. C.; Shaw J. T.; Franz A. K. Enantioselective Si-H Insertion Reactions of Diarylcarbenes for the Synthesis of Silicon-Stereogenic Silanes. J. Am. Chem. Soc. 2020, 142, 11674–11679. 10.1021/jacs.0c04533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Mu D.; Yuan W.; Chen S.; Wang N.; Yang B.; You L.; Zu B.; Yu P.; He C. Streamlined Construction of Silicon-Stereogenic Silanes by Tandem Enantioselective C-H Silylation/Alkene Hydrosilylation. J. Am. Chem. Soc. 2020, 142, 13459–13468. 10.1021/jacs.0c04863. [DOI] [PubMed] [Google Scholar]
  24. Zhang J.; Yan N.; Ju C.-W.; Zhao D. Nickel(0)-Catalyzed Asymmetric Ring Expansion Toward Enantioenriched Silicon-Stereogenic Benzosiloles. Angew. Chem., Int. Ed. 2021, 60, 25723–25728. 10.1002/anie.202111025. [DOI] [PubMed] [Google Scholar]
  25. Zhou H.; Bae H. Y.; Leutzsch M.; Kennemur J. L.; Bécart D.; List B. The Silicon–Hydrogen Exchange Reaction: A Catalytic σ-Bond Metathesis Approach to the Enantioselective Synthesis of Enol Silanes. J. Am. Chem. Soc. 2020, 142, 13695–13700. 10.1021/jacs.0c06677. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Zhou H.; Zhang P.; List B. The Silicon–Hydrogen Exchange Reaction: Catalytic Kinetic Resolution of 2-Substituted Cyclic Ketones. Synlett 2021, 32, 1953–1956. 10.1055/a-1670-5829. [DOI] [Google Scholar]
  27. Mahlau M.; List B. Asymmetric Counteranion-Directed Catalysis: Concept, Definition, and Applications. Angew. Chem., Int. Ed. 2013, 52, 518–533. 10.1002/anie.201205343. [DOI] [PubMed] [Google Scholar]
  28. Gatzenmeier T.; van Gemmeren M.; Xie Y.; Höfler D.; Leutzsch M.; List B. Asymmetric Lewis Acid Organocatalysis of the Diels–Alder Reaction by a Silylated C–H Acid. Science 2016, 351, 949–952. 10.1126/science.aae0010. [DOI] [PubMed] [Google Scholar]
  29. Zhang Z.; Bae H. Y.; Guin J.; Rabalakos C.; van Gemmeren M.; Leutzsch M.; Klussmann M.; List B. Asymmetric Counteranion-Directed Lewis Acid Organocatalysis for the Scalable Cyanosilylation of Aldehydes. Nat. Commun. 2016, 7, 12478. 10.1038/ncomms12478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Lee S.; Kaib P. S. J.; List B. Asymmetric Catalysis via Cyclic, Aliphatic Oxocarbenium Ions. J. Am. Chem. Soc. 2017, 139, 2156–2159. 10.1021/jacs.6b11993. [DOI] [PubMed] [Google Scholar]
  31. Bae H. Y.; Höfler D.; Kaib P. S. J.; Kasaplar P.; De C. K.; Döhring A.; Lee S.; Kaupmees K.; Leito I.; List B. Approaching Sub-PPM-Level Asymmetric Organocatalysis of a Highly Challenging and Scalable Carbon–Carbon Bond Forming Reaction. Nat. Chem. 2018, 10, 888. 10.1038/s41557-018-0065-0. [DOI] [PubMed] [Google Scholar]
  32. Schreyer L.; Kaib P. S. J.; Wakchaure V. N.; Obradors C.; Properzi R.; Lee S.; List B. Confined Acids Catalyze Asymmetric Single Aldolizations of Acetaldehyde Enolates. Science 2018, 362, 216–219. 10.1126/science.aau0817. [DOI] [PubMed] [Google Scholar]
  33. Schreyer L.; Properzi R.; List B. IDPi Catalysis. Angew. Chem., Int. Ed. 2019, 58, 12761–12777. 10.1002/anie.201900932. [DOI] [PubMed] [Google Scholar]
  34. Zhu C.; Mandrelli F.; Zhou H.; Maji R.; List B. Catalytic Asymmetric Synthesis of Unprotected β2-Amino Acids. J. Am. Chem. Soc. 2021, 143, 3312–3317. 10.1021/jacs.1c00249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Amatov T.; Tsuji N.; Maji R.; Schreyer L.; Zhou H.; Leutzsch M.; List B. Confinement-Controlled, Either syn- or anti-Selective Catalytic Asymmetric Mukaiyama Aldolizations of Propionaldehyde Enolsilanes. J. Am. Chem. Soc. 2021, 143, 14475–14481. 10.1021/jacs.1c07447. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Zhou H.; Zhou Y.; Bae H. Y.; Leutzsch M.; Li Y. H.; De C. K.; Cheng G.-J.; List B. Organocatalytic stereoselective cyanosilylation of small ketones. Nature 2022, 605, 84–89. 10.1038/s41586-022-04531-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Inokuma Y.; Yoshioka S.; Ariyoshi J.; Arai T.; Hitora Y.; Takada K.; Matsunaga S.; Rissanen K.; Fujita M. X-ray analysis on the nanogram to microgram scale using porous complexes. Nature 2013, 495, 461–466. 10.1038/nature11990. [DOI] [PubMed] [Google Scholar]
  38. Inokuma Y.; Yoshioka S.; Ariyoshi J.; Arai T.; Fujita M. Preparation and guest-uptake protocol for a porous complex useful for ’crystal-free’ crystallography. Nat. Protoc. 2014, 9, 246–252. 10.1038/nprot.2014.007. [DOI] [PubMed] [Google Scholar]
  39. Hoshino M.; Khutia A.; Xing H.; Inokuma Y.; Fujita M. The crystalline sponge method updated. IUCrJ. 2016, 3, 139–151. 10.1107/S2052252515024379. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Zigon N.; Duplan V.; Wada N.; Fujita M. Crystalline Sponge Method: X-ray Structure Analysis of Small Molecules by Post-Orientation within Porous Crystals—Principle and Proof-of-Concept Studies. Angew. Chem., Int. Ed. 2021, 60, 25204–25222. 10.1002/anie.202106265. [DOI] [PubMed] [Google Scholar]
  41. Li X. C.; Ferreira D.; Ding Y. Determination of absolute configuration of natural products: theoretical calculation of electronic circular dichroism as a tool. Curr. Org. Chem. 2010, 14, 1678–1697. 10.2174/138527210792927717. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Sainz M.; Souto J.; Regentova D.; Johansson M.; Timhagen S.; Irvine D. J.; Buijsen P.; Koning C.; Stockman R.; Howdle S. M. A Facile and Green Route to Terpene Derived Acrylate and Methacrylate Monomers and Simple Free Radical Polymerisation to Yield New Renewable Polymers and Coatings. Polym. Chem. 2016, 7, 2882–2887. 10.1039/C6PY00357E. [DOI] [Google Scholar]
  43. Andersen C.; Ferey V.; Daumas M.; Bernardelli P.; Guérinot A.; Cossy J. Introduction of Cyclopropyl and Cyclobutyl Ring on Alkyl Iodides through Cobalt-Catalyzed Cross-Coupling. Org. Lett. 2019, 21, 2285–2289. 10.1021/acs.orglett.9b00579. [DOI] [PubMed] [Google Scholar]
  44. Zhang G.; Li Y.; Wang Y.; Zhang Q.; Xiong T.; Zhang Q. Asymmetric Synthesis of Silicon-Stereogenic Silanes by Copper-Catalyzed Desymmetrizing Protoboration of Vinylsilanes. Angew. Chem., Int. Ed. 2020, 59, 11927–11931. 10.1002/anie.202005341. [DOI] [PubMed] [Google Scholar]
  45. Rendler S.; Oestreich M. Conclusive Evidence for an SN2-Si Mechanism in the B(C6F5)3-Catalyzed Hydrosilylation of Carbonyl Compounds: Implications for the Related Hydrogenation. Angew. Chem., Int. Ed. 2008, 47, 5997–6000. 10.1002/anie.200801675. [DOI] [PubMed] [Google Scholar]
  46. Sommer L. H.; Lyons J.; Fujimoto H. Stereochemistry of Asymmetric Silicon. XV. Stereospecific Hydrosilation and Exchange Reactions of R3Si*H(D) Catalyzed by Group VIII Metal Centers. J. Am. Chem. Soc. 1969, 91, 7051–7061. 10.1021/ja01053a027. [DOI] [Google Scholar]
  47. Sommer L.; McLick J.; Golino C. SNi-Si Mechanism. Reductive Displacement of Good Leaving Groups with Retention of Configuration by Diisobutylaluminum Hydride. Stereochemical and Mechanistic Crossover with the Etherate Complex of Diisobutylaluminum Hydride. J. Am. Chem. Soc. 1972, 94, 669–670. 10.1021/ja00757a078. [DOI] [Google Scholar]
  48. Shintani R.; Maciver E. E.; Tamakuni F.; Hayashi T. Rhodium-Catalyzed Asymmetric Synthesis of Silicon-Stereogenic Dibenzooxasilines via Enantioselective Transmetalation. J. Am. Chem. Soc. 2012, 134, 16955–16958. 10.1021/ja3076555. [DOI] [PubMed] [Google Scholar]
  49. Jeon M.; Han J.; Park J. Transformation of Silanes into Silanols Using Water and Recyclable Metal Nanoparticle Catalysts. ChemCatChem. 2012, 4, 521–524. 10.1002/cctc.201100456. [DOI] [Google Scholar]
  50. Sommer L. H.; Lyons J.-E. Stereochemistry of Asymmetric Silicon. XVI. Transition Metal Catalyzed Substitution Reactions of Optically Active Organosilicon Hydrides. J. Am. Chem. Soc. 1969, 91, 7061–7067. 10.1021/ja01053a028. [DOI] [Google Scholar]
  51. Wang Z.-C.; Wang M.; Gao J.; Shi S.-L.; Xu Y. nBuLi-Promoted anti-Markovnikov Selective Hydroboration of Unactivated Alkenes and Internal Alkynes. Org. Chem. Front. 2019, 6, 2949–2953. 10.1039/C9QO00750D. [DOI] [Google Scholar]

Associated Data

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

Supplementary Materials

ja2c04261_si_001.pdf (19.4MB, pdf)

Articles from Journal of the American Chemical Society are provided here courtesy of American Chemical Society

RESOURCES