Next Article in Journal
In Vitro Study on Anti-Hepatitis C Virus Activity of Spatholobus suberectus Dunn
Next Article in Special Issue
Theoretical Study on the Second Hyperpolarizailities of Oligomeric Systems Composed of Carbon and Silicon π-Structures
Previous Article in Journal
Novel Natural Product- and Privileged Scaffold-Based Tubulin Inhibitors Targeting the Colchicine Binding Site
Previous Article in Special Issue
Synthesis of Thioethers by InI3-Catalyzed Substitution of Siloxy Group Using Thiosilanes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reactions of an Isolable Dialkylsilylene with Aroyl Chlorides. A New Route to Aroylsilanes

Key Laboratory of Organosilicon Chemistry and Material Technology of Ministry of Education, Hangzhou Normal University, Hangzhou 311121, China
*
Authors to whom correspondence should be addressed.
Molecules 2016, 21(10), 1376; https://doi.org/10.3390/molecules21101376
Submission received: 14 September 2016 / Revised: 3 October 2016 / Accepted: 8 October 2016 / Published: 15 October 2016
(This article belongs to the Special Issue Advances in Silicon Chemistry)

Abstract

:
The reactions of isolable dialkylsilylene 1 with aromatic acyl chlorides afforded aroylsilanes 3a3c exclusively. Aroylsilanes 3a3c were characterized by 1H-, 13C-, and 29Si-NMR spectroscopy, high-resolution mass spectrometry (HRMS), and single-crystal molecular structure analysis. The reaction mechanisms are discussed in comparison with related reaction of 1 with chloroalkanes and chlorosilanes.

Graphical Abstract

1. Introduction

Acylsilanes or α-silyl ketones have been known as a unique class of silicon compounds [1,2,3,4,5,6,7,8,9], showing remarkably red–red shifted n→π* transition bands [1,2] and being useful as distinct reagents in organic synthesis [4,5,6,10,11,12,13,14,15,16,17,18,19]. Most of all, acyltris(trimethylsilyl)silanes are of particular importance, which were utilized for the synthesis of the first stable silicon–carbon doubly bonded compounds (silenes) [20,21]. However, the synthesis of acylsilanes is still limited because of the relatively facile silicon–carbon bond cleavage under the reaction conditions. The direct reaction of a silylmetal with an acyl halide afforded the corresponding acylsilanes, but the yields were usually low due to the undesired secondary reactions [22,23]. The oxidation of α-silyl alcohols using ordinary oxidizing reagents often leads to the corresponding aldehydes [5]. The first successful synthesis of an aroylsilane was achieved by using an elaborate two-step route in good yields (Equation (1)) [22], while it is not applicable for the synthesis of alkanoylsilanes.
Molecules 21 01376 i001
The dithiane route applicable for the synthesis of a wider range of acylsilanes was studied by Brook et al. [24] and Corey et al. [25] at the same time in 1967 (Equation (2)). The defect of the method is the use of a toxic mercury compound for the hydrolysis of the silylated dithianes.
Molecules 21 01376 i002
A variety of acylsilanes have been synthesized up to date using different methods, the reactions of protected aldehydes, esters, and other carboxylic acid derivatives, etc. with various silicon reagents [1,2,3,4,5,6,7,8,9].
During the course of our studies of the reactions of an isolable dialkylsilylene with various functional groups [26,27,28,29,30], we have found that the silylene inserts exclusively into the C–Cl bond of aroyl chlorides providing rather exceptional aroyl(chloro)silanes that cannot be obtained via conventional methods. Very recently, an acyl(halo)silane was utilized to synthesize an isolable silenyllithium (Equation (3)) [31].
Molecules 21 01376 i003

2. Results and Discussion

2.1. Synthesis and Characterization

The 1:1 reactions of dialkylsilylene 1 [32,33,34,35,36,37] with benzoyl and 4-substituted benzoyl chlorides 2a2c at −30 °C afforded the corresponding benzoyl(chloro)silanes 3a3c in high yields, indicating that the C(carbonyl)–Cl bond is much more reactive than the carbonyl group (Equation (4)) [38]. No significant difference was observed in the reactivity among benzoyl chlorides 2a2c. Even when an excess amount of 1 was used to a benzoyl chloride (2:1 mol ratio), the corresponding benzoyl(chloro)silane was obtained solely as the product. The expected reactions of 3 with silylene 1 would be prohibited due to the steric effects of bulky silylene moiety of 3. The reactions of 1 with alkanoyl chlorides like acetyl chloride and butanoyl chloride afforded complex reaction mixtures. Because simple alkanoyl chlorides are more reactive than aroyl chlorides, the products of the reactions between 1 and the alkanoyl chlorides may react further with 1 to give the unidentified products.
Molecules 21 01376 i004
Benzoylsilanes 3a3c, which are stable thermally with definite melting points and under moist air, were characterized by 1H-, 13C-, and 29Si-NMR spectroscopy, high-resolution mass spectrometry (HRMS), and X-ray structure analyses.

2.2. NMR Spectroscopy

In the 1H-NMR spectra of 3a3c, two singlet signals due to four trimethylsilyl groups were observed in the region of 0.1–0.3 ppm [0.16, 0.30 (3a); 0.17, 0.29 (3b); 0.17, 0.29 (3c)], indicating that there are two types of trimethylsilyl groups because of their Cs symmetry of 3. In accord with the observation, two TMS carbon (δ ca. 3.2 and 4.3 ppm) and silicon signals (δ ca. 2.8 and 5.5 ppm) were observed in the 13C- and 29Si-NMR spectra of 3a3c. The signals at 225.3 (3a), 224.5 (3b), and 224.8 (3c) ppm in 13C-NMR spectra are ascribed to the carbonyl carbon signals, which are at higher field relative to the typical acylsilanes (ca. 240 ppm) [39,40]. However, these chemical shift values are significantly lower than those for typical ketones like benzophene (δ 196.7) and acetophenone (δ 198.2), indicating the unique electronic feature of acylsilanes. The 29Si-NMR resonances due to the ring silicon of 3a3c appear at the same chemical shifts of 27.8 ppm.

2.3. Molecular Structure Analysis

Molecular structures of compounds 3a3c were determined by X-ray single-crystal diffraction analysis. Yellow single crystals of 3a3c suitable for X-ray crystallography were obtained by slowly evaporating the solvent from their hexane solutions. The ORTEP drawing of compound 3a is depicted in Figure 1. Compound 3a was crystallized in space group P-1 with two crystallographically independent molecules in an asymmetric unit. The structural parameters of the two molecules in a unit cell are similar but different in the torsion angles of C(1)Si(1)C(17)O(1) and its equivalent, C(24)Si(6)C(40)O(2), (129.03° and 5.26°, respectively). The sum of bond angles around C(17) and C(40) are 360°, being in accord with the sp2 character of the carbonyl carbon atom. The distances of Si(1)–C(17) (1.935(3) Å) and Si(6)–C(40) bonds (1.929(2) Å), are significantly larger than the normal Si–C bond length (1.87–1.89 Å). A similarly long distance of the Si–C(carbonyl) bond (1.926 Å) has been observed in the molecular structure of acetyltriphenylsilane by Trotter et al. [41]. The origin may be ascribed to the effective σ(SiC)–n(O) conjugation as proposed by Ramsey, Brook, Bassindale, and Bock [42]. In other words, it is suggested that resonance form B contributed significantly to the bonding in acylsilanes.
Similarly, compounds 3b and 3c were crystallized in space group P21/n and P-1 and their molecular structures are shown in Figure 2 and Figure 3. A single crystal of 3b has two crystallographically independent molecules in the asymmetric unit, while that of 3c has one independent molecule. Their structural parameters are similar to those of 3a.

2.4. Mechanistic Aspects

Because the insertion of silylene 1 into C–Cl [43,44,45,46,47,48] and Si–Cl [48,49,50,51,52,53] bonds have been reported, it would be desirable to propose the mechanisms of the present acylsilane formation as being consistent with the features of these precedents. The reactions of isolable dialkylsilylene 1 with chloroalkanes afford rather unusual product mixtures depending on the substrates. For example, 1 reacts with 1-chlorobutane to afford solely the corresponding butylchlorosilane 4, while the reaction of 1 with CCl4 gives only dichlorosilane 5 (Equation (5)) [43]. When cyclopropylmethyl chloride is used as a substrate, the rather unusual 2:1 adduct 6 was obtained in addition to 5 (Equation (6)) [43].
Molecules 21 01376 i005
Molecules 21 01376 i006
The diverse modes of the reactions between 1 with chloroalkanes [43] suggest a complex nature of the mechanisms. The reactions may be understood uniformly starting from initially formed Lewis acid-base complexes as shown in Scheme 1. From the complex, ionic cleavage of the C–Cl bond followed by recombination would yield an alkylchlorosilane such as 4 [43]. The ionic mechanism is also applicable for the reaction of 1 with cyclopropylmethyl chloride, in which the intermediary cyclopropylmethyl cation or its equivalent 3-butenyl cation reacts with an extra silylene 1 forming 3-butenylsilyl cation and then finally 6; the 3-butenylsilyl cation would be stabilized by the coordination of the terminal π bond. Chloroalkanes with less electron donating substituents like CHCl3 and CCl4 destabilize the carbocation intermediates and instead yield 5 after the homolysis of the C–Cl bond [54].
The aroylation of 1 may occur concertedly from the acylsilane-silylene complex as shown in Equation (7). Alternatively, the facile heterolysis of the C(carbonyl)–Cl bond from the complex followed by the coupling in cage may occur exclusively; the silylene serves as a Lewis acid to activate the C(carbonyl)–Cl bond (Equation (7)). The former concerted mechanism is preferred to the latter because of the similarity of the reactions with those of chlorislanes with 1 [50,51].
Molecules 21 01376 i007
The insertion reactions of silylene 1 into the Si–Cl bonds of chlorosilanes have been found to occur cleanly [49,50]; hence, the concerted mechanism via three-membered cyclic transition states has been proposed. The mechanism has been supported by the detailed DFT calculations [55,56,57].
Molecules 21 01376 i008

3. Materials and Methods

3.1. General Procedures

Manipulation of air-sensitive compounds was performed under a controlled dry argon atmosphere using standard Schlenk techniques. Tetrahydrofuran (THF), hexane, and toluene were distilled from sodium–benzophenone. All the other reagents were obtained from commercial suppliers and used without further purification. Dialkylsilylene 1 was prepared according to literature procedures [32]. 1H- (400 MHz), 13C- (100.6 MHz), and 29Si- (79.5 MHz) NMR spectra were recorded on a Bruker AV-400 spectrometer at room temperature (Bruker, Rheinstetten, Germany), using CDCl3 as the solvent. Melting points are uncorrected. High-resolution mass spectra (HRMS) were recorded on a Bruker Daltonics Apex-III spectrometer (Bruker, Rheinstetten, Germany).

3.2. Synthesis

3.2.1. Synthesis of 3a

A hexane solution of benzoyl chloride (0.45 g, 3.2 mmol) was added to a solution of dialkylsilylene 1 (1.12 g, 3.0 mmol) in hexane at −30 °C. The reaction mixture was allowed to stir for 2 h at 0 °C. The color of the solution changed from red to yellow. Then, the solvent was removed under vacuum. The resulting residue was purified by flash chromatography (Silica gel, 200–300 mesh; ethyl acetate/hexane, 1:300) to yield 3a as a yellow solid. Yield: 0.98 g (64%). m.p. 152–154 °C; 1H-NMR (400 MHz, CDCl3): δ 8.07 (d, J = 7.2 Hz, 2H, o-Ar), 7.56 (t, J = 7.2 Hz, 1H, p-Ar), 7.49 (t, J = 7.2 Hz, 2H, m-Ar), 2.14 (s, 4H, CH2), 0.30 (s, 18H, SiMe3), 0.16 (s, 18H, SiMe3). 13C-NMR (101 MHz, CDCl3): δ 225.28 (C=O),138.90 (CAr-C(O)),133.22 (p-Ar), 129.42 (o-Ar), 128.30 (m-Ar), 33.25 (CH2), 12.50 (C(SiMe3)2), 4.27 (SiMe3), 3.25 (SiMe3); 29Si-NMR (80 MHz, CDCl3): δ 27.84 (SiCl), 5.53 (SiMe3), 2.85 (SiMe3); HRMS(ESI) calculated for C23H45ClOSi5: 513.2079, found 513.2078.

3.2.2. Synthesis of 3b

A hexane solution of p-methyl benzoyl chloride (0.49 g, 3.2 mmol) was added to a solution of dialkylsilylene 1 (1.12 g, 3.0 mmol) in hexane at −30 °C. The reaction mixture was allowed to stir for 2 h at 0 °C. The color of the solution changed from red to yellow. Then, the solvent was removed under vacuum. The resulting residue was purified by flash chromatography (Silica gel, 200–300 mesh; ethyl acetate/hexane, 1:300) to yield 3b as a yellow solid. Yield: 0.97 g (61%). m.p. 174–177 °C; 1H-NMR (400 MHz, CDCl3): δ 7.96 (d, J = 7.2 Hz, 2H, o-Ar), 7.29 (d, J = 7.2 Hz, 2H, m-Ar), 2.43 (s, 3H, Ar-Me), 2.13 (s, 4H, CH2), 0.29 (s, 18H, SiMe3), 0.17 (s, 18H, SiMe3). 13C-NMR (101 MHz, CDCl3): δ 224.53 (C=O), 144.13 (CAr-C(O)), 150.00, 136.63 (p-Ar), 129.59 (o-Ar), 128.98 (m-Ar), 33.24 (CH2), 21.77 (Ar-Me), 12.44 (C(SiMe3)2), 4.28 (SiMe3), 3.24 (SiMe3). 29Si-NMR (79 MHz, CDCl3): δ 27.84 (SiCl), 5.47 (SiMe3), 2.81 (SiMe3); HRMS(ESI) calculated for C24H47ClOSi5: 527.2225, found 527.2235.

3.2.3. Synthesis of 3c

A hexane solution of p-trifluoromethyl benzoyl chloride (0.67 g, 3.2 mmol) was added to a solution of dialkylsilylene 1 (1.12 g, 3.0 mmol) in hexane at −30 °C. The reaction mixture was allowed to stir for 2 h at 0 °C. The color of the solution changed from red to yellow. Then, the solvent was removed under vacuum. The resulting residue was purified by flash chromatography (Silica gel, 200–300 mesh; ethyl acetate/hexane, 1:300) to yield 3c as a yellow solid. Yield: 1.16 g, (67%). m.p. 160–163 °C; 1H-NMR (400 MHz, CDCl3): δ 8.17 (d, J = 7.2 Hz, 2H, o-Ar), 7.76 (d, J = 7.2 Hz, 2H, m-Ar), 2.16 (s, 4H, CH2), 0.29 (s, 18H, SiMe3), 0.17 (s, 18H, SiMe3). 13C-NMR (101 MHz, CDCl3): δ 224.78 (C=O), 140.87 (CAr-C(O)), 134.31 (dd, p-Ar), 129.57 (m-Ar), 124.86 (o-Ar), 123.61 (dd, JC-F = 271, CF3), 33.26 (CH2), 12.61 (C(SiMe3)2), 4.29 (SiMe3), 3.22 (SiMe3). 29Si-NMR (80 MHz, CDCl3) δ 27.83 (SiCl), 5.66 (SiMe3), 2.94 (SiMe3); HRMS(ESI) calculated for C24H44ClF3OSi5: 581.1949, found 581.1952.

3.3. X-ray Crystallography

The diffraction data of 3a3c were collected on a Bruker Smart Apex II CCD diffractometer with graphite-monochromated Mo Kα radiation (λ = 0.71073 Å). All of the data were collected at ambient temperatures, and the structures were solved via the direct method and subsequently refined on F2 using full-matrix least-squares techniques (SHELXTL) [58]. Absorption corrections were applied empirically using the SADABS program [59]. The non-hydrogen atoms were refined anisotropically, and hydrogen atoms were located at calculated positions. A summary of the crystallographic data and selected experimental information is given in Table S1.

4. Conclusions

Isolable dialkylsilylene 1 was found to react with the C(carbonyl)–Cl bonds in aroyl chlorides 2 at low temperatures highly chemoselectively to give aroyl(chloro)silanes 3; the carbonyl groups in neither 2 nor 3 react with silylene 1. The structural analysis using NMR and X-ray crystallography indicate the lower field 13C-NMR resonance of the carbonyl carbon and longer Si–C(carbonyl) bond distance than the standard values. The facile and highly selective nature of the reactions suggests that the insertion occurs concertedly from the initial Lewis acid-base complexes, similarly to that of 1 into the Si–Cl bonds in chlorosilanes. We are hoping the present synthetic methodology is applicable in general for a wide variety of silylenes. The silylenes should be however relatively long-lived and their reactions with the aroyl chlorides should be fast enough to prevent their oligomerization. Further works on the acylsilanes with unique electronic properties are under progress in our laboratory.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/21/10/1376/s1. Crystallographic information for compounds 3a3c in CIF format and crystallographic tables.

Acknowledgments

This work was financially supported by the National Natural Science Foundation of China (Grant No. 21472032) and the Natural Science Foundation of Zhejiang Province (Grant No. LY17B010002).

Author Contributions

Z.L. conceived and designed the experiments; X.L., Q.L. and X.-Q.X. performed the experiments; X.-Q.X., X.L., Q.L., Z.L., G.L. and M.K. analyzed the data; X.-Q.X., Z.L. and M.K. wrote the paper. All authors have given approval to the final version of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Crystallography data (excluding structure factors) for the structures reported in this paper have been deposited with the Cambridge Crystallographic Data Center, CCDC 979677 (3a), 979676 (3b) and 979675 (3c). Copies of these data can be obtained free of charge on application to the Director, CCDC, 12 Union Road, Cambridge CB2 1EZ, UK (fax: +44-1223-336033; e-mail: [email protected] or http:// www.http.ccdc.cam.ac.hk).

References and Notes

  1. Brook, A.G. Keto Derivatives of Group IV Organometalloids. In Advances in Organometallic Chemistry; Stone, F.G.A., Robert, W., Eds.; Academic Press: Cambridge, MA, USA, 1969; Volume 7, pp. 95–155. [Google Scholar]
  2. Page, P.C.B.; Klair, S.S.; Rosenthal, S. Synthesis and chemistry of acyl silanes. Chem. Soc. Rev. 1990, 19, 147–195. [Google Scholar] [CrossRef]
  3. Page, P.C.B.; Mckenzie, M.J.; Klair, S.S.; Rosenthal, S. Acyl Silanes. In The Chemistry of Organic Silicon Compounds; Rappoport, Z., Apeloig, Y., Eds.; John Wiley & Sons: Hoboken, NJ, USA, 1998; pp. 1599–1665. [Google Scholar]
  4. Cirillo, P.F.; Panek, J.S. Recent progress in the chemistry of acylsilanes. A review. Org. Prep. Proced. Int. 1992, 24, 553–582. [Google Scholar] [CrossRef]
  5. Patrocínio, A.F.; Moran, P.J.; Brazil, J. Acylsilanes and their applications in organic chemistry. Chem. Soc. 2001, 12, 7–31. [Google Scholar] [CrossRef]
  6. Zhang, H.-J.; Priebbenow, D.L.; Bolm, C. Acylsilanes: valuable organosilicon reagents in organic synthesis. Chem. Soc. Rev. 2013, 42, 8540–8571. [Google Scholar] [CrossRef] [PubMed]
  7. Page, P.C.B.; McKenzie, M.J. Product Subclass 25: Acylsilanes. In Science of Synthesis; Fleming, I., Ed.; Thieme: Stuttgart, Germany, 2001; Volume 4, pp. 513–568. [Google Scholar]
  8. Garrett, M.N.; Johnson, J.S. Product Subclass 4: Silicon Compounds. In Science of Synthesis Knowledge Updates 2012/2; Fleming, I., Ed.; Thieme: Stuttgart, Germany, 2012; Volume 4, pp. 1–85. [Google Scholar]
  9. Boyce, G.R.; Grezler, S.N.; Johnson, J.S.; Linghu, X.; malinovski, J.T.; Nicewicz, D.A.; Satterfield, A.D.; Schmitt, D.C.; Steward, K.M. Silyl Glyoxylates. Conception and Realization of Flexible Conjunctive Reagents for Multicomponent Coupling. J. Org. Chem. 2012, 77, 4503–4515. [Google Scholar] [CrossRef] [PubMed]
  10. Tsubouchi, A.; Sasaki, N.; Enatsu, S.; Takeda, T. Regio-and stereoselective preparation of (Z)-silyl enol ethers by three-component coupling using α, β-unsaturated acylsilanes as core building blocks. Tetrahedron Lett. 2013, 54, 1264–1267. [Google Scholar] [CrossRef]
  11. Honda, M.; Nakajima, T.; Okada, M.; Yamaguchi, K.; Suda, M.; Kunimoto, K.K.; Segi, M. Reaction of acylsilanes with α-sulfinyl carbanions: Regioselective synthesis of silyl enol ethers. Tetrahedron Lett. 2011, 52, 3740–3742. [Google Scholar] [CrossRef]
  12. Honda, M.; Iwamoto, R.; Nogami, Y.; Segi, M. Stereoselective tandem Aldol-Tishchenko reaction with acylsilanes. Chem. Lett. 2005, 34, 466–467. [Google Scholar] [CrossRef]
  13. Zhang, H.J.; Becker, P.; Huang, H.; Pirwerdjan, R.; Pan, F.F.; Bolm, C. Photochemically induced silylacylations of alkynes with acylsilanes. Adv. Synth. Catal. 2012, 354, 2157–2161. [Google Scholar] [CrossRef]
  14. Wu, L.; Li, G.X.; Fu, Q.Q.; Yu, L.T.; Tang, Z. Organocatalytic asymmetric Michael reaction with acylsilane donors. Org. Biomol. Chem. 2013, 11, 443–447. [Google Scholar] [CrossRef] [PubMed]
  15. Chen, J.P.; Ding, C.H.; Liu, W.; Hou, X.L.; Dai, L.X. Palladium-catalyzed regio-, diastereo-, and enantioselective allylic alkylation of acylsilanes with monosubstituted allyl substrates. J. Am. Chem. Soc. 2010, 132, 15493–15495. [Google Scholar] [CrossRef] [PubMed]
  16. Xin, L.H.; Nicewicz, D.A.; Johnson, J.S. Tandem Carbon−Carbon Bond Constructions via Catalyzed Cyanation/Brook Rearrangement/C-Acylation Reactions of Acylsilanes. Org. Lett. 2002, 4, 2957–2960. [Google Scholar]
  17. Mattson, A.E.; Scheidt, K.A. Catalytic Additions of Acylsilanes to Imines: An Acyl Anion Strategy for the Direct Synthesis of α-Amino Ketones. Org. Lett. 2004, 6, 4363–4366. [Google Scholar] [CrossRef] [PubMed]
  18. Ruiz, J.; Karre, N.; Roisnel, T.; Chandrasekhar, S.; Gree, R. From Protected β-Hydroxy Acylsilanes to Functionalized Silyl Enol Ethers and Applications in Mukaiyama Aldol Reactions. Eur. J. Org. Chem. 2016, 773–779. [Google Scholar] [CrossRef]
  19. Murthy, A.S.; Roisnel, T.; Chandrasekhar, S.; Gree, R. New β-Hydroxy Acylsilane-Derived Building Blocks and Their Use in the Synthesis of Oxygen-Containing Heterocycles. Synlett 2013, 24, 2216–2220. [Google Scholar] [CrossRef]
  20. Brook, A.G.; Abdesaken, F.; Gutekunst, B.; Gutekunst, G.; Kallury, R.K. A Solid Silaethene: Isolation and Characterization. J. Chem. Soc. Chem. Commun. 1981, 191–192. [Google Scholar] [CrossRef]
  21. Brook, A.G.; Nyburg, S.C.; Abdesaken, F.; Gutekunst, B.; Gutekunst, G.; Krishna, R.; Kallury, M.R.; Poon, Y.C.; Chang, Y.-M.; Wong-Ng, W. Stable Solid Silaethylenes. J. Am. Chem. Soc. 1982, 104, 5667–5612. [Google Scholar] [CrossRef]
  22. Brook, A.G. Triphenylsilyl phenyl ketone. J. Am. Chem. Soc. 1957, 79, 4373–4375. [Google Scholar] [CrossRef]
  23. Wittenberg, D.; Gilman, H. Reactions of Silyllithium Compounds with Derivatives of Carboxylic Acids. I. Triphenylsilyllithium and Acetyl Chloride. J. Am. Chem. Soc. 1958, 80, 4529–4531. [Google Scholar] [CrossRef]
  24. Brook, A.G.; Duff, J.M.; Jones, P.F.; Davis, N.R. Synthesis of Silyl and Germyl Ketones. J. Am. Chem. Soc. 1967, 89, 431–434. [Google Scholar] [CrossRef]
  25. Corey, E.J.; Seebach, D.; Freedman, R. Synthesis of α-Silyl Ketones via 1,3-Dithianes. J. Am. Chem. Soc. 1967, 89, 434–436. [Google Scholar] [CrossRef]
  26. Chen, W.F.; Wang, L.L.; Li, Z.F.; Lin, A.Q.; Lai, G.Q.; Xiao, X.Q.; Deng, Y.; Kira, M. Diverse reactivity of an isolable dialkylsilylene toward imines. Dalton Trans. 2013, 42, 1872–1878. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, X.P.; Xiao, X.Q.; Xu, Z.; Yang, X.M.; Li, Z.F.; Dong, Z.W.; Yan, C.T.; Lai, G.Q.; Kira, M. Reactions of an Isolable Dialkylsilylene with Carbon Dioxide and Related Heterocumulenes. Organometallics 2014, 33, 5434–5439. [Google Scholar] [CrossRef]
  28. Wang, L.L.; Chen, W.F.; Li, Z.F.; Xiao, X.Q.; Lai, G.Q.; Liu, X.P.; Xu, Z.; Kira, M. Reactions of an Isolable Dialkylsilylene with Aromatic Nitriles Providing a New Type of Heterosilole. Chem. Commun. 2013, 49, 9776–9778. [Google Scholar] [CrossRef] [PubMed]
  29. Dong, Z.; Xiao, X.Q.; Li, Z.; Lu, Q.; Lai, G.; Kira, M. Elusive 2H-1,2-Oxasiletes Through Reactions of an Isolable Dialkylsilylene with Diazocarbonyl Compounds. Org. Biomol. Chem. 2015, 13, 9471–9476. [Google Scholar] [CrossRef] [PubMed]
  30. Xiao, X.Q.; Dong, Z.; Li, Z.; Yan, C.; Lai, G.; Kira, M. 1,3-Diazasilabicyclo[1.1.0]butane with a Long Bridging N−N Bond. Angew. Chem. Int. Ed. 2016, 55, 3758–3762. [Google Scholar] [CrossRef] [PubMed]
  31. Pinchuk, D.; Mathew, J.; Kaushansky, A.; Bravo-Zhivotovskii, D.; Apeloig, Y. Isolation and Characterization, Including by X-ray Crystallography, of Contact and Solvent-Separated Ion Pairs of Silenyl Lithium Species. Angew. Chem. Int. Ed. 2016, 55, 10258–10262. [Google Scholar] [CrossRef] [PubMed]
  32. Kira, M.; Ishida, S.; Iwamoto, T.; Kabuto, C. The first isolable dialkylsilylene. J. Am. Chem. Soc. 1999, 121, 9722–9723. [Google Scholar] [CrossRef]
  33. Kira, M. Isolable silylene, disilenes, trisilaallene, and related compounds. J. Organomet. Chem. 2004, 689, 4475–4488. [Google Scholar] [CrossRef]
  34. Kira, M.; Ishida, S.; Iwamoto, T. Comparative chemistry of isolable divalent compounds of silicon, germanium, and tin. Chem. Rec. 2004, 4, 243–253. [Google Scholar] [CrossRef] [PubMed]
  35. Kira, M.; Iwamoto, T.; Ishida, S. A Helmeted Dialkylsilylene. Bull. Chem. Soc. Jpn. 2007, 80, 258–275. [Google Scholar] [CrossRef]
  36. Kira, M. An isolable dialkylsilylene and its derivatives. A step toward comprehension of heavy unsaturated bonds. Chem. Commun. 2010, 46, 2893–2903. [Google Scholar] [CrossRef] [PubMed]
  37. Kira, M. Reactions of a Stable Dialkylsilylene and Their Mechanisms. J. Chem. Sci. 2012, 124, 1205–1215. [Google Scholar] [CrossRef]
  38. Ishida, S.; Iwamoto, T.; Kira, M. Reactions of an Isolable Dialkylsilylene with Ketones. Organometallics 2010, 29, 5526–5534. [Google Scholar] [CrossRef] and references cited therein.
  39. Ohshita, J.; Tokunaga, Y.; Sakurai, H.; Kunai, A. Reactions of lithium silenolates with acyl halides. First synthesis of di-and tetraacylsilanes. J. Am. Chem. Soc. 1999, 121, 6080–6081. [Google Scholar] [CrossRef]
  40. Ohshita, J.; Kawamoto, H.; Kunai, A.; Ottosson, H. Formation of Acylsilenolates from Bis (acyl) trisilanes as the Silicon Analogues of Acylenolates. Organometallics 2010, 29, 4199–4202. [Google Scholar] [CrossRef]
  41. Chieh, P.; Trotter, J. The structure of acetyltriphenylsilane. J. Chem. Soc. A 1969, 1778–1783. [Google Scholar] [CrossRef]
  42. Ramsey, B.G.; Brook, A.; Bassindale, A.R.; Bock, H. σ→π*, A reassignment of the long wavelength uv transition in acyl-silanes and-germanes by photoelectron spectroscopy. J. Organomet. Chem. 1974, 74, C41–C45. [Google Scholar] [CrossRef]
  43. Ishida, S.; Iwamoto, T.; Kabuto, C.; Kira, M. Unexpected reactions of an isolable dialkylsilylene with haloalkanes. Chem. Lett. 2001, 1102–1103. [Google Scholar] [CrossRef]
  44. Ishikawa, M.; Nakagawa, K.-I.; Katayama, S.; Kumada, M. Photolysis of organopolysilanes. The reaction of photochemically generated trimethylsilyphenylsilylene with alkyl chlorides. J. Organomet. Chem. 1981, 216, C48–C50. [Google Scholar] [CrossRef]
  45. Nakao, R.; Oka, K.; Dohmaru, T.; Nagata, Y.; Fukumoto, T. Chlorine abstraction from chloromethanes by dimethylsilanediyls. J. Chem. Soc. Chem. Commun. 1985, 766–768. [Google Scholar] [CrossRef]
  46. Oka, K.; Nakao, R. Reaction of phenyl (trimethylsily) silylene with chloromethanes; insertion into the C-Cl bond and abstraction of chlorine and HCI. J. Organomet. Chem. 1990, 390, 7–18. [Google Scholar] [CrossRef]
  47. Moser, D.F.; Bosse, T.; Olson, J.; Moser, J.L.; Guzei, I.A.; West, R. Halophilic Reactions of a Stable Silylene with Chloro and Bromocarbons. J. Am. Chem. Soc. 2002, 124, 4186–4187. [Google Scholar] [CrossRef] [PubMed]
  48. Xiong, Y.; Yao, S.; Driess, M. Reactivity of a Zwitterionic Stable Silylene toward Halosilanes and Haloalkanes. Organometallics 2009, 28, 1927–1933. [Google Scholar] [CrossRef]
  49. Ishida, S.; Iwamoto, T.; Kabuto, C.; Kira, M. A stable silicon-based allene analogue with a formally sp-hybridized silicon atom. Nature 2003, 421, 725–727. [Google Scholar] [CrossRef] [PubMed]
  50. Ishida, S.; Iwamoto, T.; Kabuto, C.; Kira, M. Insertion of a stable dialkylsilylene into silicon-chlorine bonds. Silicon Chem. 2003, 2, 137–140. [Google Scholar] [CrossRef]
  51. Chen, Y.S.; Gaspar, P.P. Octakis(triemthylsilyl)cyclotetrasilane. A stable cyclotetrasilane from a silylene precursor. Organometallics 1982, 1, 1410–1412. [Google Scholar] [CrossRef]
  52. Belzner, J.; Dehnert, U.; Ihmels, H.; Hübner, M.; Müller, P.; Uson, I. New Dichlorosilanes, Cyclotrisilanes, and Silacyclopropanes as Precursors of Intramolecularly Coordinated Silylenes. Chem. Eur. J. 1998, 4, 852–863. [Google Scholar] [CrossRef]
  53. Belzner, J.; Dehnert, U.; Schar, D.; Rohde, B.; Muller, P.; Uson, I. Synthesis of di- and trisilanes with potentially chelating substituents. J. Organomet. Chem. 2002, 649, 25–42. [Google Scholar] [CrossRef]
  54. Koecher, J.; Lehnig, M.; Neumann, W.P. Chemistry of heavy carbene analogs R2M (M = Si, Ge, Sn). 12. Concerted and nonconcerted insertion reactions of dimethylgermylene into the carbon-halogen bond. Organometallics 1988, 7, 1201–1207. [Google Scholar] [CrossRef]
  55. Lai, G.; Xu, Z.; Li, Z.; Jiang, J.; Kira, M.; Qiu, H. Stereoelectronic Substituent Effects on Silylene Insertion into the Si-Cl Bond. Organometallics 2009, 28, 3591–3593. [Google Scholar] [CrossRef]
  56. Xu, Z.; Jin, J.; Li, Z.; Qiu, H.; Jiang, J.; Lai, G.; Kira, M. Remarkable Substituent Effects on the Activation Energy of Silylene Insertion into Silicon-Chlorine Bonds. Chem. Eur. J. 2009, 15, 8605–8612. [Google Scholar] [CrossRef] [PubMed]
  57. Xu, Z.; Jin, J.; Zhang, H.; Li, Z.; Jiang, J.; Lai, G.; Kira, M. Insertion of Silylenes into Si-H and Si-Cl Bonds. Comparison of Mechanism and Substituent Effects. Organometallics 2011, 30, 3311–3317. [Google Scholar] [CrossRef]
  58. SHELXTL version 6. 10; Bruker AXS Inc.: Madison, WI, USA, 2003.
  59. Sheldrick, G.M. SADABS Program for Empirical X-ray Absorption Correction; University of Goettingen: Göttingen, Germany, 1996. [Google Scholar]
  • Sample Availability: Samples of the compounds 3a3c are available from the authors.
Figure 1. ORTEP drawing of the independent molecules in 3a. (Hydrogen atoms are omitted for clarity. Thermal ellipsoids are shown at the 30% probability level.) Selected bond lengths (Å) and angles (°): Si(1)–C(4) 1.877(2), Si(1)–C(1) 1.890(2), Si(1)–C(17) 1.935(3), Si(1)–Cl(1) 2.1006(9), O(1)–C(17) 1.223(3), Si(6)–C(24) 1.882(2), Si(6)–C(27) 1.889(2), Si(6)–C(40) 1.929(2), Si(6)–Cl(2) 2.1001(9), O(2)–C(40) 1.227(3); C(4)–Si(1)–C(1) 102.55(11), C(17)–Si(1)–Cl(1) 95.66(9), O(1)–C(17)–C(18) 120.2(2), O(1)–C(17)–Si(1) 114.54(19), C(18)–C(17)–Si(1) 125.21(17), C(24)–Si(6)–C(27) 102.25(11), C(40)–Si(6)–Cl(2) 96.14(8), O(2)–C(40)–C(41) 119.4(2), O(2)–C(40)–Si(6) 114.64(18), C(41)–C(40)–Si(6) 125.89(17).
Figure 1. ORTEP drawing of the independent molecules in 3a. (Hydrogen atoms are omitted for clarity. Thermal ellipsoids are shown at the 30% probability level.) Selected bond lengths (Å) and angles (°): Si(1)–C(4) 1.877(2), Si(1)–C(1) 1.890(2), Si(1)–C(17) 1.935(3), Si(1)–Cl(1) 2.1006(9), O(1)–C(17) 1.223(3), Si(6)–C(24) 1.882(2), Si(6)–C(27) 1.889(2), Si(6)–C(40) 1.929(2), Si(6)–Cl(2) 2.1001(9), O(2)–C(40) 1.227(3); C(4)–Si(1)–C(1) 102.55(11), C(17)–Si(1)–Cl(1) 95.66(9), O(1)–C(17)–C(18) 120.2(2), O(1)–C(17)–Si(1) 114.54(19), C(18)–C(17)–Si(1) 125.21(17), C(24)–Si(6)–C(27) 102.25(11), C(40)–Si(6)–Cl(2) 96.14(8), O(2)–C(40)–C(41) 119.4(2), O(2)–C(40)–Si(6) 114.64(18), C(41)–C(40)–Si(6) 125.89(17).
Molecules 21 01376 g001
Figure 2. ORTEP drawing of the independent molecules in 3b. (Hydrogen atoms are omitted for clarity. Thermal ellipsoids are shown at the 30% probability level.) Selected bond lengths (Å) and angles (°): Si(1)–C(1) 1.883(4), Si(1)–C(4) 1.898(4), Si(1)–C(17) 1.940(4), Si(1)–Cl(1) 2.1003(14), C(17)–O(1) 1.223(5), Si(6)–C(28) 1.888(4), Si(6)–C(25) 1.887(4), Si(6)–C(41) 1.927(4), Si(6)–Cl(2) 2.0967(15), C(41)–O(2) 1.216(5); C(1)–Si(1)–C(4) 102.26(17), C(17)–Si(1)–Cl(1) 97.97(13), O(1)–C(17)–C(18) 119.7(4), O(1)–C(17)–Si(1) 113.2(3), C(18)–C(17)–Si(1) 126.9(3), C(28)–Si(6)–C(25) 102.12(16), C(41)–Si(6)–Cl(2) 98.33(14), O(2)–C(41)–C(42) 119.1(4), O(2)–C(41)–Si(6) 114.4(3), C(42)–C(41)–Si(6) 126.3(3).
Figure 2. ORTEP drawing of the independent molecules in 3b. (Hydrogen atoms are omitted for clarity. Thermal ellipsoids are shown at the 30% probability level.) Selected bond lengths (Å) and angles (°): Si(1)–C(1) 1.883(4), Si(1)–C(4) 1.898(4), Si(1)–C(17) 1.940(4), Si(1)–Cl(1) 2.1003(14), C(17)–O(1) 1.223(5), Si(6)–C(28) 1.888(4), Si(6)–C(25) 1.887(4), Si(6)–C(41) 1.927(4), Si(6)–Cl(2) 2.0967(15), C(41)–O(2) 1.216(5); C(1)–Si(1)–C(4) 102.26(17), C(17)–Si(1)–Cl(1) 97.97(13), O(1)–C(17)–C(18) 119.7(4), O(1)–C(17)–Si(1) 113.2(3), C(18)–C(17)–Si(1) 126.9(3), C(28)–Si(6)–C(25) 102.12(16), C(41)–Si(6)–Cl(2) 98.33(14), O(2)–C(41)–C(42) 119.1(4), O(2)–C(41)–Si(6) 114.4(3), C(42)–C(41)–Si(6) 126.3(3).
Molecules 21 01376 g002
Figure 3. ORTEP drawing of compound 3c. (Hydrogen atoms are omitted for clarity. Thermal ellipsoids are shown at the 30% probability level.) Selected bond lengths (Å) and angles (°): Si(1)–C(1) 1.8789(17), Si(1)–C(4) 1.8850(16), Si(1)–C(17) 1.9364(19), Si(1)–Cl(1) 2.1029(7), C(17)–O(1) 1.217(2); C(1)–Si(1)–C(4) 102.20(7), C(17)–Si(1)–Cl(1) 7.03(6), O(1)–C(17)–C(18) 18.54(16), O(1)–C(17)–Si(1) 15.19(14), C(18)–C(17)–Si(1) 126.03(13).
Figure 3. ORTEP drawing of compound 3c. (Hydrogen atoms are omitted for clarity. Thermal ellipsoids are shown at the 30% probability level.) Selected bond lengths (Å) and angles (°): Si(1)–C(1) 1.8789(17), Si(1)–C(4) 1.8850(16), Si(1)–C(17) 1.9364(19), Si(1)–Cl(1) 2.1029(7), C(17)–O(1) 1.217(2); C(1)–Si(1)–C(4) 102.20(7), C(17)–Si(1)–Cl(1) 7.03(6), O(1)–C(17)–C(18) 18.54(16), O(1)–C(17)–Si(1) 15.19(14), C(18)–C(17)–Si(1) 126.03(13).
Molecules 21 01376 g003
Scheme 1. Mechanisms of the reactions of 1 with chloroalkanes.
Scheme 1. Mechanisms of the reactions of 1 with chloroalkanes.
Molecules 21 01376 sch001

Share and Cite

MDPI and ACS Style

Xiao, X.-Q.; Liu, X.; Lu, Q.; Li, Z.; Lai, G.; Kira, M. Reactions of an Isolable Dialkylsilylene with Aroyl Chlorides. A New Route to Aroylsilanes. Molecules 2016, 21, 1376. https://doi.org/10.3390/molecules21101376

AMA Style

Xiao X-Q, Liu X, Lu Q, Li Z, Lai G, Kira M. Reactions of an Isolable Dialkylsilylene with Aroyl Chlorides. A New Route to Aroylsilanes. Molecules. 2016; 21(10):1376. https://doi.org/10.3390/molecules21101376

Chicago/Turabian Style

Xiao, Xu-Qiong, Xupeng Liu, Qiong Lu, Zhifang Li, Guoqiao Lai, and Mitsuo Kira. 2016. "Reactions of an Isolable Dialkylsilylene with Aroyl Chlorides. A New Route to Aroylsilanes" Molecules 21, no. 10: 1376. https://doi.org/10.3390/molecules21101376

Article Metrics

Back to TopTop