Abstract
Chiral organosilanes do not exist in nature and are therefore absent from the “chiral pool”. As a consequence, synthetic approaches toward enantiopure silanes, stereogenic at silicon, are rather limited. While catalytic asymmetric desymmetrization reactions of symmetric organosilicon compounds have been developed, the utilization of racemic silanes in a dynamic kinetic asymmetric transformation (DYKAT) or dynamic kinetic resolution (DKR) would significantly expand the breadth of accessible Si-stereogenic compounds. We now report a DYKAT of racemic allyl silanes enabled by strong and confined imidodiphosphorimidate (IDPi) catalysts, providing access to Si-stereogenic silyl ethers. The products of this reaction are easily converted into useful enantiopure monohydrosilanes. We propose a spectroscopically and experimentally supported mechanism involving the epimerization of a catalyst-bound intermediate.
In light of their importance in chemical synthesis and technical materials, enantiopure silicon-containing compounds bearing a Si-stereogenic center have recently gained considerable attention.1−10 However, while C-stereogenic congeners are readily accessible via a plethora of asymmetric transformations of C=X π-bonds (X = C, O, N), the synthesis of silicon-stereogenic silanes has been notoriously difficult due to the instability and therefore unavailability of the corresponding Si=X-based molecules.11,12 Documented enantioselective approaches to Si-stereogenic organosilanes can be divided into four main categories: chromatographic resolution,13 chemical resolution with a stoichiometric chiral reagent,14 catalytic desymmetrization of symmetric silicon compounds,3−5,7,15−17 and catalytic kinetic resolutions (eq 1).18−21 We became interested in contributing potentially more general and higher-yielding synthetic strategies toward enantiopure organosilanes. Specifically, we anticipated to expand the synthetic toolbox by efficiently utilizing racemic silanes, either in a dynamic kinetic resolution (DKR) or a dynamic kinetic asymmetric transformation (DYKAT), which both have the inherent advantage of offering yields >50%. A rhodium-catalyzed dynamic kinetic asymmetric hydrosilylation has recently been described by Xu and co-workers, but organocatalysis has not previously been used in a DYKAT toward stereogenic silanes.22
![]() |
1 |
However, since preservation of stereochemical integrity is frequently observed in organosilicon chemistry, a challenge to overcome in such strategies is the required racemization of the Si-stereogenic center during the reaction conditions.23−26 Inspiration came from previous work on the desymmetrization of symmetrical silanes via a silicon hydrogen exchange reaction and general studies on silylium-based asymmetric counteranion-directed catalysis (Si-ACDC).17,27−50 Accordingly, we focused our attention on the design of a strong and confined imidodiphosphorimidate (IDPi)-catalyzed asymmetric synthesis of Si-stereogenic silyl ethers from racemic allyl silanes (eq 2). We anticipated that this approach could benefit from the configurational lability of the silylated IDPi intermediate, enabling high enantioselectivity and full conversion of the racemic starting material. Here we report the fruition of these considerations with an IDPi-catalyzed DYKAT of racemic allyl silanes with phenols, providing Si-stereogenic silyl ethers in excellent yield and enantioselectivity.
After identifying methallyl silane 1a and propofol (2) as suitable reactants in an initial screening (see Tables S1–S3 in the Supporting Information), we began optimizing this model reaction as depicted in Table 1.51 First, different solvents were examined with IDPi catalyst 3a (entries 1–4), and toluene was identified as a preferred solvent, for example, compared to methylcyclohexane, in which scarce solubility of the IDPi catalyst was observed. We also explored variations of the aryl group on the catalyst. Indeed, IDPi catalyst 3b, bearing p-tert-butyl phenyl groups at the 3,3′-positions of the binaphthyl backbone and triflyl groups at the inner core, enabled the formation of the desired product in 94:6 e.r. (entry 5). Excitingly, the e.r. of product 4a could be further improved to 99:1 by evolving the inner core (entries 6 and 7). It is noteworthy that increasing the reaction concentration suppressed the formation of side products, leading to satisfactory yields with no significant effect on the enantioselectivity (entries 8 and 9).
Table 1. Reaction Development.
Unless stated otherwise, reactions were performed with rac-1a (0.025 mmol), propofol 2 (1.5 equiv), and IDPi catalysts 3a–d (2.5 mol %) in solvent (0.25 mL, 0.1 M).
Estimated by 1H NMR or TLC analysis.
Determined by 1H NMR using CH2Br2 as internal standard.
Enantiomeric ratios (e.r.) determined by HPLC.
0.2 M.
0.4 M.
Under the optimized conditions, the scope and generality of the IDPi-catalyzed DYKAT of silanes was evaluated and the results are outlined in Table 2. Chiral silyl ether 4a was obtained in 70% yield with 99:1 e.r. when the reaction of rac-1a and phenol 2 was conducted on a 0.2 mmol scale. Similarly, racemic silanes bearing different linear alkyl chains could all be accommodated to provide the corresponding products 4b–f with good to excellent results (67–84% yields and 85:15 to 99:1 e.r.). Additionally, a branched-alkyl-substituted silane was also efficiently converted, furnishing the corresponding product 4g in 95% yield and 98:2 e.r.. We found that high levels of enantioinduction were maintained when silane starting materials bearing substituents with diverse electronic properties in para and meta positions on the phenyl group were subjected to the reaction conditions. Organosilanes 4h–l were readily obtained with excellent results (77–87% yields and 96.5:3.5 to 99:1 e.r.). In contrast, a moderate e.r. was obtained from the reaction of silane 1m, which bears an ortho substituent. 2-Naphthyl and thienyl moieties were compatible with the DYKAT process, affording products 4n,o in good yields and excellent enantioselectivities. In addition to being competent at the synthesis of organosilanes bearing saturated alkyl chains, a range of racemic silanes featuring alkenyl groups were tolerated equally well, enabling access to products 4p–r in 61–76% yield and 86:14 to 97:3 e.r. However, the enantiodifferentiation was less effective in the presence of a benzyl-substituted racemic silane, and product 4s was obtained in 70% yield and 75:25 e.r. It is noteworthy that the reactivity of the starting silanes is highly dependent on the nature of the allyl leaving groups; in fact, the lack of reactivity of internal alkenyl and simple allyl moieties correlates to their nucleophilicity.52 Moreover, we were curious to investigate substrates with an existing carbon stereocenter in the deracemization reaction with propofol. The transformation of a 1:1 diastereomeric mixture of the enantiopure starting material 1v proved to be exquisitely catalyst controlled, and the two enantiomers of IDPi 3d gave products 4t,u with 18:1 and >20:1 d.r., respectively. Furthermore, our reaction is readily scalable and a preparative synthesis of Si-stereogenic silyl ether 4a (1.8 g) was accomplished in 82% isolated yield without erosion of enantioselectivity and the catalyst was recovered in 88% yield. More importantly, monohydrosilane 5 could easily be obtained without significant loss of enantiopurity by the reductive removal of the phenolic unit with retention of Si-stereochemistry.53 The described two-step approach here may facilitate the wide utilization of such hydrosilanes as chiral auxiliaries, protecting groups, reagents, and synthetic precursors.54
Table 2. Substrate Scopea.
Unless specified otherwise, reactions were performed with rac-1 (0.2 mmol, 1.0 equiv), 2,6-diisopropylphenol 2 (0.3 mmol, 1.5 equiv), and IDPi 3d (2.5 mol %) in toluene (0.1 M) at −20 °C for 4–7 days. Isolated yields are given, and enantiomeric ratios (e.r.) were determined by HPLC. Conditions of reactions starting from a silicon compound bearing a C-stereocenter: 1v (0.1 mmol, 1.0 equiv), 2 (0.15 mmol, 1.5 equiv), and IDPi 3d (2.5 mol %) in toluene (0.1 M) at −20 °C for 7 days. Conditions for gram-scale synthesis and product elaboration: rac-1 (5.8 mmol, 1.0 equiv), 2 (8.7 mmol, 1.5 equiv), and IDPi 3d (2.5 mol %) in toluene (0.1 M) at −20 °C for 3 days; 4a (0.3 mmol) and DIBAL-H (3 equiv) in hexanes (0.15 M) at rt for 3 days.
–40 °C.
0.4 M.
0.2 M.
rt.
With IDPi 3b as catalyst.
With 5 mol % catalyst.
To shed light on the underlying reaction mechanism of the discovered DYKAT, we conducted a time-course study for the reaction of allyl silane 1a with propofol 2 catalyzed by IDPi 3d in toluene at −20 °C for 3 days. The enantiopurity of the remaining silane 1a and product 4a was continuously monitored during the course of the reaction (Figure 1A). Interestingly, the IDPi catalyst enables nearly complete control of the enantioselectivity of product 4a from the starting point of the reaction, while the remaining silane 1a undergoes a slow enantioenrichment throughout the course of the reaction, suggesting a kinetic resolution process. Meanwhile, racemization studies under IDPi catalysis were performed with enantiopure starting materials (S)-1a and (R)-1a, respectively, and no racemization was observed in both cases, excluding a DKR pathway (Figure 1B). Further, we investigated the activation of catalyst 3d with rac-1a by 31P NMR analysis. Upon addition of racemic silane 1a (40 equiv), the sharp singlet of IDPi 3d (1.0 equiv) present at −4.5 ppm (Figure 1C, (i)) broadened due to the averaging of the acquired spectra during the drying of trace amounts of adventitious water (Figure 1C, (ii)). This signal sharpened again and shifted significantly to −6.3 ppm after complete drying (Figure 1C, (iii)). Finally, the catalyst was fully silylated, presumably mainly into diastereomers Cat-Si1 and Cat-Si2 and two additional minor species. The two main diastereomers feature two major doublets (Jpp = 97.4 Hz) at −2.0 and −7.0 ppm and two minor doublets (Jpp = 95.1 Hz) at −3.6 and −6.9 ppm, reflecting covalent bonding between the silicon atom and presumably an oxygen atom of the catalyst (Figure 1C, (iv)).49 It is interesting to note that there is interconversion between the epimers, which was confirmed by comparing the 31P NMR spectrum of the catalyst with those of (R)-1a, (S)-1a ,and rac-1a (see Figures S5–S10 in the Supporting Information). We also separately investigated the reaction of the three starting materials (R)-1a, (S)-1a, and rac-1a under the standard conditions. The same product enantiomer (S)-4a was obtained in all three cases, with similar results, showing that the IDPi catalyst largely controls the stereoselectivity, irrespective of enantiopurity and absolute configuration of the starting silane (see Figure S1 in the Supporting Information). This indicates the existence of similar intermediates, which likely are positively charged, as validated by a Hammett plot analysis (see Figure S12 in the Supporting Information). Furthermore, we monitored the reversion of the silylated catalyst into the protonated form and found the Si-(R) enantiomer to react much faster than its Si-(S) counterpart (Figure 1E,F).
Figure 1.
Mechanistic studies. (A) Time-course study with rac-1a (0.1 mmol, 1.0 equiv), 2 (0.15 mmol, 1.5 equiv), and IDPi 3d (2.5 mol %) in toluene (0.1 M) at −20 °C under Ar for 3 days. Conversions were determined by GC, and enantiomeric ratios (e.r.) were determined by HPLC. (B) Racemization studies of either (R)-1a or (S)-1a (0.1 mmol) and IDPi 3d (2.5 mol %) in toluene (0.2 M) at −20 °C for 3 days. (C) 31P NMR of the activation of catalyst with rac-1a (0.1 mmol, 1.0 equiv) and IDPi 3d (2.5 mol %) in toluene-d8 (0.1 M) at −20 °C under Ar. (D) Silylation process performed under the same conditions as in (C), monitored by 31P NMR. (E) Reaction of (S)-1a, performed with (S)-1a (0.1 mmol, 1.0 equiv), 2 (0.15 mmol, 1.5 equiv), and IDPi 3d (2.5 mol %) in toluene-d8 (0.1 M) at −20 °C under Ar. The variations of Cat-Si1 and Cat-H were monitored by 31P NMR. (F) Reaction of (R)-1a under the same conditions as in (E).
Computational studies at the PBE-D3/def2-TZVP+CPCM (toluene) level of theory (see the Supporting Information for details) indicate that the diastereomers feature similar structures and a negligible energy difference (0.4 kcal/mol), which is within the expected computational error. Indeed, Cat-Si1 and Cat-Si2 show Si–O bond lengths of 1.81 and 1.82 Å, respectively. Nonetheless, they interconvert readily under the experimental conditions, with a reaction barrier of 19.2 kcal/mol, which is consistent with the experimental findings (Figure 2A, left). In order to explain the mode of interconversion, we located the transition state; this structure suggests the existence of a pentacoordinated silicon species, in which the Si atom simultaneously establishes two Si–O bonds with comparable lengths (2.22 and 2.16 Å). As a result, the silicon fragment and the two oxygen atoms of the IDPi’s sulfonyl groups arrange in space with a distorted-bipyramidal-trigonal geometry (Figure 2A, right).
Figure 2.
(A) Computational studies and (B) proposed mechanism.
Based on the accumulated experimental and computational results, a mechanism of IDPi-catalyzed DYKAT of racemic silane can be proposed (Figure 2B). First, the reaction sequence of protonation/C–Si bond cleavage of the silane enantiomers takes place, furnishing active silyl-catalyst species Cat-Si1 and Cat-Si2, which interconvert as a result of the silyl group migration between the two diastereotopic oxygen atoms in the active site of the catalyst. The S enantiomer of the product is then generated through nucleophilic attack on Cat-Si2 that is more favorable over the reaction of Cat-Si1.
In summary, we have developed an organocatalytic DYKAT of racemic allyl silanes using propofol as a nucleophile. A variety of functionalized silicon-stereogenic silyl ethers were constructed and readily derivatized to monohydrosilanes, potentially serving as general building blocks for chemical synthesis. Moreover, we elucidated a possible epimerization route triggered by the IDPi catalyst and supported a proposed mechanism by experimental and theoretical studies. By providing a general and direct transformation of racemic silanes into optically active Si-stereogenic compounds with high efficiency, this study paves a new avenue for organosilicon chemistry, which we believe will stimulate numerous applications in polymer and materials science, as well as in medicinal chemistry.
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 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.3c00858.
Experimental details and analytical data for all new compounds (PDF)
Author Contributions
H.Z. and R.P. contributed equally.
Open access funded by Max Planck Society.
The authors declare the following competing financial interest(s): a patent on the general catalyst class has been filed.
Supplementary Material
References
- Hiyama T.; Shirakawa E. Organosilicon compounds. Cross-Coupling Reactions 2002, 219, 61–85. 10.1007/3-540-45313-X_3. [DOI] [Google Scholar]
- 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]
- 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]
- 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]
- Shintani R. Recent Advances in the Transition-Metal-Catalyzed Enantioselective Synthesis of Silicon-Stereogenic Organosilanes. Asian J. Org. Chem. 2015, 4, 510–514. 10.1002/ajoc.201500066. [DOI] [Google Scholar]
- Cui Y. M.; Lin Y.; Xu L. W. Catalytic synthesis of chiral organoheteroatom compounds of silicon, phosphorus, and sulfur via asymmetric transition metal-catalyzed C-H functionalization. Coord. Chem. Rev. 2017, 330, 37–52. 10.1016/j.ccr.2016.09.011. [DOI] [Google Scholar]
- Shintani R. Recent progress in catalytic enantioselective desymmetrization of prochiral organosilanes for the synthesis of silicon-stereogenic compounds. Synlett 2018, 29, 388–396. 10.1055/s-0036-1591839. [DOI] [Google Scholar]
- Hiyama T.; Oestreich M.. Organosilicon Chemistry: Novel Approaches and Reactions; Wiley: 2020. [Google Scholar]
- Zheng L.; Nie X. X.; Wu Y.; Wang P. Construction of Si-Stereogenic Silanes through C-H Activation Approach. Eur. J. Org. Chem. 2021, 2021, 6006–6014. 10.1002/ejoc.202101084. [DOI] [Google Scholar]
- He C.; Yuan W. Enantioselective C-H Functionalization toward Silicon-Stereogenic Silanes. Synthesis 2022, 54, 1939–1950. 10.1055/a-1729-9664. [DOI] [Google Scholar]
- Ahlrichs R.; Heinzmann R. Stability and reactivity of the silicon-carbon double bond. J. Am. Chem. Soc. 1977, 99, 7452–7456. 10.1021/ja00465a009. [DOI] [Google Scholar]
- Morkin T. L.; Leigh W. J. Substituent Effects on the Reactivity of the Silicon-Carbon Double Bond. Acc. Chem. Res. 2001, 34, 129–136. 10.1021/ar960252y. [DOI] [PubMed] [Google Scholar]
- Mori A.; Toriyama F.; Kajiro H.; Hirabayashi K.; Nishihara Y.; Hiyama T. Synthesis and optical resolution of novel chiral silanols. Chem. Lett. 1999, 28, 549–550. 10.1246/cl.1999.549. [DOI] [Google Scholar]
- Sommer L.; Frye C.; Parker G.; Michael K.-W. Stereochemistry of Asymmetric Silicon. I. Relative and Absolute Configurations of Optically Active α-Naphthylphenylmethylsilanes. J. Am. Chem. Soc. 1964, 86, 3271–3276. 10.1021/ja01070a014. [DOI] [Google Scholar]
- Bauer J. O.; Strohmann C. Recent progress in asymmetric synthesis and application of difunctionalized silicon-stereogenic silanes. Eur. J. Inorg. Chem. 2016, 2016, 2868–2881. 10.1002/ejic.201600100. [DOI] [Google Scholar]
- 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]
- Zhou H.; Han J. T.; Nöthling N.; Lindner M. M.; Jenniches J.; Kühn C.; Tsuji N.; Zhang L.; List B. Organocatalytic Asymmetric Synthesis of Si-Stereogenic Silyl Ethers. J. Am. Chem. Soc. 2022, 144, 10156–10161. 10.1021/jacs.2c04261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamamoto K.; Kawanami Y.; Miyazawa M. Kinetic resolution of (±)-cyclohex-1-enylsilanols by the sharpless asymmetric epoxidation. J. Chem. Soc., Chem. 1993, 5, 436–437. 10.1039/C39930000436. [DOI] [Google Scholar]
- Tacke R.; Heinrich T. Syntheses of enantiopure Si-centrochiral silicon-based muscarinic antagonists using an enantioselective enzymatic esterification as the key step. Silicon Chem. 2002, 1, 35–39. 10.1023/A:1016095507813. [DOI] [Google Scholar]
- Rendler S.; Auer G.; Keller M.; Oestreich M. Preparation of a Privileged Silicon-Stereogenic Silane: Classical versus Kinetic Resolution. Adv. Synth. Catal. 2006, 348, 1171–1182. 10.1002/adsc.200606071. [DOI] [Google Scholar]
- Guo Y.; Liu M. M.; Zhu X.; Zhu L.; He C. Catalytic Asymmetric Synthesis of Silicon-Stereogenic Dihydrodibenzosilines: Silicon Central-to-Axial Chirality Relay. Angew. Chem., Int. Ed. 2021, 133, 14006–14010. 10.1002/ange.202103748. [DOI] [PubMed] [Google Scholar]
- During the preparation of the manuscript, a Rh-catalyzed DYKAT of hydrosilanes was reported:; Zeng Y.; Fang X. J.; Tang R. H.; Xie J. Y.; Zhang F. J.; Xu Z.; Nie Y. X.; Xu L. W. Rhodium-Catalyzed Dynamic Kinetic Asymmetric Hydrosilylation to Access Silicon-Stereogenic Center. Angew. Chem., Int. Ed. 2022, 61, e202214147 10.1002/anie.202214147. [DOI] [PubMed] [Google Scholar]
- Bauer J. O.; Strohmann C. From an α-Functionalized Silicon-Stereogenic N, O-Silane to a Monomeric and Tetracoordinate tBuLi Adduct with Lithium-Centered Chirality. Angew. Chem., Int. Ed. 2014, 53, 8167–8171. 10.1002/anie.201404255. [DOI] [PubMed] [Google Scholar]
- Igawa K.; Kokan N.; Tomooka K. Asymmetric Synthesis of Chiral Silacarboxylic Acids and Their Ester Derivatives. Angew. Chem., Int. Ed. 2010, 49, 728–731. 10.1002/anie.200904922. [DOI] [PubMed] [Google Scholar]
- 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]
- Ducos P.; Liautard V.; Robert F.; Landais Y. Chiral Memory in Silylium Ions. Chem. -Eur. J. 2015, 21, 11573–11578. 10.1002/chem.201501987. [DOI] [PubMed] [Google Scholar]
- 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]
- Wang Q.; Leutzsch M.; van Gemmeren M.; List B. Disulfonimide-catalyzed asymmetric synthesis of β3-amino esters directly from N-Boc-Amino sulfones. J. Am. Chem. Soc. 2013, 135, 15334–15337. 10.1021/ja408747m. [DOI] [PubMed] [Google Scholar]
- Wang Q.; van Gemmeren M.; List B. Asymmetric disulfonimide-catalyzed synthesis of δ-amino-β-ketoester derivatives by vinylogous Mukaiyama-Mannich reactions. Angew. Chem., Int. Ed. 2014, 53, 13592–13595. 10.1002/anie.201407532. [DOI] [PubMed] [Google Scholar]
- 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]
- Kaib P. S.; Schreyer L.; Lee S.; Properzi R.; List B. Extremely active organocatalysts enable a highly enantioselective addition of allyltrimethylsilane to aldehydes. Angew. Chem., Int. Ed. 2016, 55, 13200–13203. 10.1002/anie.201607828. [DOI] [PubMed] [Google Scholar]
- Lee S.; Kaib P. S.; 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]
- Bae H. Y.; Höfler D.; Kaib P. S.; 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]
- Gatzenmeier T.; Turberg M.; Yepes D.; Xie Y.; Neese F.; Bistoni G.; List B. Scalable and Highly Diastereo-and Enantioselective Catalytic Diels-Alder Reaction of α, β-Unsaturated Methyl Esters. J. Am. Chem. Soc. 2018, 140, 12671–12676. 10.1021/jacs.8b07092. [DOI] [PubMed] [Google Scholar]
- Schreyer L.; Kaib P. S.; 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]
- Bae H. Y.; List B. Triflimide: An Overlooked High-Performance Catalyst of the Mukaiyama Aldol Reaction of Silyl Ketene Acetals with Ketones. Eur. J. Chem. 2018, 24, 13767–13772. 10.1002/chem.201803142. [DOI] [PubMed] [Google Scholar]
- Lee S.; Bae H. Y.; List B. Can a ketone be more reactive than an aldehyde? Catalytic asymmetric synthesis of substituted tetrahydrofurans. Angew. Chem., Int. Ed. 2018, 57, 12162–12166. 10.1002/anie.201806312. [DOI] [PubMed] [Google Scholar]
- Schreyer L.; Properzi R.; List B. IDPi Catalysis. Angew. Chem., Int. Ed. 2019, 58, 12761–12777. 10.1002/anie.201900932. [DOI] [PubMed] [Google Scholar]
- Mandrelli F.; Blond A.; James T.; Kim H.; List B. Deracemizing α-Branched Carboxylic Acids by Catalytic Asymmetric Protonation of Bis-Silyl Ketene Acetals with Water or Methanol. Angew. Chem., Int. Ed. 2019, 58, 11479–11482. 10.1002/anie.201905623. [DOI] [PubMed] [Google Scholar]
- Zhang Z.; Klussmann M.; List B. Kinetic study of disulfonimide-catalyzed cyanosilylation of aldehydes by using a method of progress rates. Synlett 2020, 31, 1593–1597. 10.1055/s-0040-1707129. [DOI] [Google Scholar]
- 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]
- 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]
- 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]
- 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]
- Das S.; Mitschke B.; De C. K.; Harden I.; Bistoni G.; List B. Harnessing the ambiphilicity of silyl nitronates in a catalytic asymmetric approach to aliphatic β3-amino acids. Nat. Catal. 2021, 4, 1043–1049. 10.1038/s41929-021-00714-x. [DOI] [Google Scholar]
- Obradors C.; List B. Azine activation via silylium catalysis. J. Am. Chem. Soc. 2021, 143, 6817–6822. 10.1021/jacs.1c03257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ouyang J.; Bae H.; Jordi S.; Dao Q. M.; Dossenbach S.; Dehn S.; Lingnau J. B.; De C. K.; Kraft P.; List B. The smelling principle of vetiver oil, unveiled by chemical synthesis. Angew. Chem., Int. Ed. 2021, 60, 5666–5672. 10.1002/anie.202014609. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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]
- Ouyang J.; Maji R.; Leutzsch M.; Mitschke B.; List B. Design of an Organocatalytic Asymmetric (4 + 3) Cycloaddition of 2-Indolylalcohols with Dienolsilanes. J. Am. Chem. Soc. 2022, 144, 8460–8466. 10.1021/jacs.2c02216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grossmann O.; Maji R.; Aukland M. H.; Lee S.; List B. Catalytic Asymmetric Additions of Enol Silanes to In Situ Generated Cyclic, Aliphatic N-Acyliminium Ions. Angew. Chem., Int. Ed. 2022, 61, e202115036 10.1002/anie.202115036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smaller o-substituents (i.e., methyl) on the phenol gave lower enantioselectivity, while larger substituents (i.e., t-Bu) gave similar reactivity. Considering practicality, we chose readily available and inexpensive propofol throughout our studies.
- Phan T. B.; Breugst M.; Mayr H. Towards a General Scale of Nucleophilicity?. Angew. Chem., Int. Ed. 2006, 45, 3869–3874. 10.1002/anie.200600542. [DOI] [PubMed] [Google Scholar]; The full Mayr’s database is available online at: https://www.cup.lmu.de/oc/mayr/reaktionsdatenbank2.
- 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]
- Wu Y.; Wang P. Silicon-Stereogenic Monohydrosilane: Synthesis and Applications. Angew. Chem., Int. Ed. 2022, 61, e202205382 10.1002/anie.202205382. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.