Next Article in Journal
In Vitro Delivery and Controlled Release of Doxorubicin for Targeting Osteosarcoma Bone Cancer
Previous Article in Journal
Ensemble-Based Interpretations of NMR Structural Data to Describe Protein Internal Dynamics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and NMR-Study of 1-Trimethylsilyl Substituted Silole Anion [Ph4C4Si(SiMe3)]•[Li]+ and 3-Silolenide 2,5-carbodianions {[Ph4C4Si(n-Bu)2]2•2[Li]+, [Ph4C4Si(t-Bu)2]2•2[Li]+} via Silole Dianion [Ph4C4Si]2•2[Li]+

Department of Nanopolymer Material Engineering, Pai Chai University, 155-40 Baejae-ro (Doma-Dong), Seo-Gu, Daejon 302-735, Korea
Dedicated to Professor Wan-Chul Joo on the occasion of his 89th birthday.
Molecules 2013, 18(9), 10568-10579; https://doi.org/10.3390/molecules180910568
Submission received: 24 July 2013 / Revised: 21 August 2013 / Accepted: 22 August 2013 / Published: 30 August 2013
(This article belongs to the Section Molecular Diversity)

Abstract

:
1-Trimethylsilyl, 1-R (R = Me, Et, i-Bu)-2,3,4,5-tetraphenyl-1-silacyclopentadiene [Ph4C4Si(SiMe3)R] are synthesized from the reaction of 1-trimethylsilyl,1-lithio-2,3,4,5-tetraphenyl-1-silacyclopentadienide anion [Ph4C4SiMe3]•[Li]+ (3) with methyl iodide, ethyl iodide, and i-butyl bromide. The versatile intermediate 3 is prepared by hemisilylation of the silole dianion [Ph4C4Si]2•2[Li]+ (2) with trimethylsilyl chloride and characterized by 1H-, 13C-, and 29Si-NMR spectroscopy. 1,1-bis(R)-2,3,4,5-tetraphenyl-1-silacyclopentadiene [Ph4C4SiR2] {R = n-Bu (7); t-Bu (8)} are synthesized from the reaction of 2 with n-butyl bromide and t-butyl bromide. Reduction of 7 and 8 with lithium under sonication gives the respective 3-silolenide 2,5-carbodianions {[Ph4C4Si(n-Bu)2]2•2[Li]+ (10) and [Ph4C4Si(t-Bu)2]2•2[Li]+ (11)}, which are characterized by 1H-, 13C-, and 29Si-NMR spectroscopy. Polarization of phenyl groups in 3 is compared with those of silole anion/dianion, germole anion/dianion, and 3-silolenide 2,5-carbodianions 10 and 11.

1. Introduction

Cyclopentadienyl anion, the most representative aromatic compound, has for a long time played important roles in organic and organometallic chemistry [1,2,3]. Therefore it has been a challenge to synthesize the analogue framework [4,5,6,7], in which one of carbon atoms is replaced by a heavier group 14 atom, and the ultimate question is to find out how its aromaticity changes and is maintained [8,9,10,11,12,13,14,15]. Since the first silacyclopentadienide dianion was reported [16], the aromaticity of sila- and germa-cyclopentadienide dianion has been suggested by NMR chemical shift changes upon reduction [17,18]. Their aromatic structures [19,20] and the related structures have been confirmed by X-ray crystallography [21,22,23,24,25,26,27]. Heavier metallic dianion equivalents, the stannacyclopentadienide dianion [28,29,30] and plumbacyclopentadienide dianion [31], are also reported to display aromaticity [32,33,34,35,36,37].
The principal heavier congener of the cyclopentadienide anion, 1-tert-butyl-2,3,4,5-tetraphenyl-1-silacyclopentadienide anion, has been reported to have aromaticity according to NMR chemical shift changes upon reduction [38]. Meanwhile 1-methyl-2,3,4,5-tetraphenyl-1-silacyclopentadienide anion was synthesized and crystallized in THF as a [2+2] dimer of its Si = C bond in aromatic ring structures [39], the dimer of which is dissociated to the original silole anions when it is reacted with alkyl halides or trimethylchlorosilane in THF [40]. Even the analogue frameworks of trimetallic anion {[C2GeSi2], [C2Si3]} and divalent germanium containing anion {[C3NGe:]}, in which more than one carbon atom of the cyclopentadienyl anion are replaced by heavier group 14 atoms of Si and/or Ge, are synthesized and characterized to have aromaticity [41,42,43], making it possible to form heavy analogues of ferrocene with them [44,45].
In contrast spectroscopic and X-ray crystallographic data [22,23] for 1-trimethylsilyl-tetramethyl/ethyl-1-silacyclopentadienide anions have revealed that they possess pyramidal silicon centers and bond localization in their butadiene moieties. Nevertheless the heavy analogues of ferrocene are synthesized with them [24,27,46,47]. Therefore it is interesting to study 1-trimethylsilyl-2,3,4,5-tetraphenyl-1-silacyclopentadienide anion [Ph4C4Si(SiMe3)] to compare it with other metallole anions [48].
There are several routes for silole syntheses, via 1,4-dilithio-butadienides by using diphenylacetylene [16,17,49] and 1,4-dihalobutadienes [23,50,51], the intramolecular reductive cyclization of diethynylsilanes [52,53], metallacyclic transfer reactions [54], and organoboration [55,56,57,58]. However those synthetic methods are not applicable to synthesizing of various siloles derivatives at the Si atom, especially for preparing 1-trimethylsilyl group substituted siloles due to the feasibility of the nucleophilic attack on the Si-Si bond by carbanions [22,24,27,59,60] and silole anion [39]. Coversely all metallole dianions are potential and useful intermediates for the synthesis of various di-substituted metallole derivatives, polysiloles, and silole-containing polymers [61,62,63,64,65,66].
Herein we report that silole dianion is a versatile intermediate to synthesize [Ph4C4Si(SiMe3)(R)] (R = Me, Et, i-Bu) via [Ph4C4Si(SiMe3)], which is prepared by hemisilylation of the silole dianion and characterized by 1H-, 13C-, and 29Si-NMR spectroscopy, and [Ph4C4SiR2] (R = i-Bu, t-Bu).

2. Results and Discussion

2.1. Preparation of 1-Trimethylsilyl,1-lithio-2,3,4,5-tetraphenyl-1-silacyclopentadienide Anion (3) and Its Reaction with Methyl Iodide, Ethyl Iodide and i-Butyl Bromide

1-Trimethylsilyl-2,3,4,5-tetraphenyl-1-silacyclopentadienide anion [Ph4C4Si(SiMe3)]•[Li]+ (3) was prepared from the reaction of the silole dianion [Ph4C4Si]−2•2[Li]+ (2) with one equivalent of trimethylsilyl chloride. The silole dianion 2 was generated by the sonication of 1,1-dichloro-2,3,4,5-tetraphenyl-1-silacyclopentadiene [Ph4C4SiCl2] (1) with lithium in THF [17]. Compound 3 in THF was reacted with the alkyl halides of methyl iodide, ethyl iodide, and i-butyl bromide to provide [Ph4C4Si(SiMe3)(R)] [R = Me (4) , Et (5), i-Bu (6)], respectively (Scheme 1).
Scheme 1. Synthesis and alkylation of 3 via silole dianion 2.
Scheme 1. Synthesis and alkylation of 3 via silole dianion 2.
Molecules 18 10568 g001
Silylation of 2 with trimethylsilyl chloride is a novel reaction to synthesize 1-trimethylsilyl substituted silole anion 3; the similar alkylation of stannole dianion with t-butyl chloride was reported to give 1-t-butyl substituted stannole anion, oxidation of which in the air provided 1,1-bis(1-t-butyl-stannole) [67,68]. Oxidation of stannole dianion were reported to give bistannole-1,2-dianion or terstannole-1,3-dianion [69,70]. But the silole anion 3 decomposes in the air to give the ring opening products of 1,2,3,4-tetraphenylbutadiene and silicate. Preparation of 4 is interesting since addition of trimethylsilyl chloride to the silole anion [Ph4C4Si(Me)]•[M]+ (M = Li, Na) in THF has given 1,1-bi(1-methyl-silole) [Ph4C4Si(Me)]2, but addition of the silole dianion to trimethylsilyl chloride in THF has provided [Ph4C4Si(SiMe3)(Me)] (4) [40].

2.2. Synthesis of 1,1-Bis(n-butyl/t-butyl)-2,3,4,5-tetraphenyl-1-silacyclopentadiene and NMR-Study of 3-Silolenide-2,5-carbodianions

1,1-bis(n-Butyl/t-butyl)-2,3,4,5-tetraphenyl-1-silacyclopentadiene {[Ph4C4Si(n-Bu)2] (7) and [Ph4C4Si(t-Bu)2] (8)} are prepared in good yield from the reactions of silole dianion 2, which is generated by the sonication of 1 with lithium in THF, with n-bromobutane and t-butyl bromide. In the case of t-butyl bromide, [Ph4C4Si(t-Bu)2] (8) is produced along with 1,1-bi[(t-Bu)SiC4Ph4] (9) in the ratio of 3 to 1 (Scheme 2). Until now there is one report of the synthesis of 1,1-bis(t-butyl)-substituted silole, which has been prepared photochemically in low yield [71].
1,1-bis(n-Butyl/t-butyl)-2,3,4,5-tetraphenyl-1-silacyclopentadiene {[Ph4C4Si(n-Bu)2] (7) and [Ph4C4Si(t-Bu)2] (8)} are sonicated in THF-d8 with lithium in the 5 mm NMR tube for 2 h. During this time the color of the mixture becomes red and/or purple. The NMR study of the reduced species in THF-d8 shows clearly that the only one species is formed and is assigned to the respective reduced 3-silolenes with 2,5-carbodianions {[Ph4C4Si(n-Bu)2]−2•2[Li]+ (10) and [Ph4C4Si(t-Bu)2]−2•2[Li]+ (11)}. Each of their 13C-NMR spectra presents ten peaks, consistent with C2 symmetry, and the 29Si spectrum of each compound shows only one resonance. The respective 1H-NMR spectrum of 10 and 11 shows two kinds of protons, 20 phenyl protons and 18 butyl protons. Even if they are sonicated further, they show the same peaks with no change.
Scheme 2. Synthesis of 7 and 8 and their reduction to 10 and 11.
Scheme 2. Synthesis of 7 and 8 and their reduction to 10 and 11.
Molecules 18 10568 g002
Both chemical shifts of the two tert-C groups (73.18 ppm for (10), 78.12 ppm for (11)) are consistent with those of the reported 3-silolenides with 2,5-carbodianions, [Ph4C4Si(R1)(R2)]−2 (77.4 ppm for R1 = R2 = Me [72], 76.42 ppm for R1 = Me, R2 = H [73]), and 1,1-R1,R2-2-lithio-2,3,4,5-tetraphenyl-1-silacyclopenta-3-enide anion (77.78 ppm for R1 = R2 = H [74]). The 13C-NMR chemical shifts of two Ciα, Ciβ and two Cpα, Cpβ show at 151.51, 147.68 ppm and 108.49, 120.80 ppm for 10 and at 152.59, 147.36 ppm and 110.87, 120.25 ppm for 11 (Table 1). The localized carbanions polarize the phenyl groups more than those of the aromatic silole/germole dianions and silole anions. The extent of polarization [Sum(Ci-Cp)/2] in those species shows in narrow range: 3-silolenides 2,5-carbodianions (34.42 to 35.00 ppm), silole/germole dianions [Ph4C4E]−2 [E = Si (2), Ge] (28.60 to 28.64 ppm), and silole anion [Ph4C4Si(t-Bu)] (24.65 ppm). In case of the phenyl group on germanium atom in the localized germole anion [Me4C4GePh]•[Li]+ [75], the extent of polarization {[Sum(Ci-Cp)/2] = 35.3 ppm} is very close to those of the localized 3-silolenide 2,5-carbodianions (Table 1 and Table 2).
Upon lithiation of 8 to 11 the 29Si-NMR chemical shift of 11 is not changed much (16.49 ppm (8) to 13.69 ppm (11) since there is no change of its hybridization with the same substituents on the silicon atom (Table 1).

2.3. NMR Study of 1-Trimethylsilyl,1-lithio-2,3,4,5-tetraphenyl-1-silacyclopentadienide Anion (3)

[Ph4C4SiCl2] (1) is sonicated in THF-d8 with lithium in a 5 mm NMR tube for 2 h, whereby the color of the mixture becomes red and/or purple. NMR study of the species in THF-d8 clearly indicates that only one species of silole dianion [Ph4C4Si]−2•2[Li]+ (2) is generated. Upon adding one equivalent of trimethylsilyl chloride to 2 the 29Si-NMR chemical shift changes from 68.54 ppm (for 2) to −13.22 ppm with another new resonance peak of the trimethylsilyl group at −15.54 ppm {[Ph4C4Si(SiMe3)]•[Li]+ (3)}. The 13C-NMR spectrum of 3 shows ten peaks in the aromatic region, consistent with C2 symmetry, and one peak for the trimethylsilyl group (Table 1). In its 1H-NMR spectrum of it there are two kinds of protons, 20 protons corresponding to four phenyl groups and 9 protons of one trimethylsilyl group.
Table 1. 13C/29Si-NMR chemical shifts of the localized 3-silolenides and germole anions.
Table 1. 13C/29Si-NMR chemical shifts of the localized 3-silolenides and germole anions.
3-Silenes 2,5-carbanion[Ph4C4SiMe2]−2•2[Li]+[Ph4C4SiMeH]−2•2[Li]+[Ph4C4Si( n-Bu)2]−2•2[Li]+ (10)[Ph4C4Si( t-Bu)2]−2 •2[Li]+ (11)[Me4C4GePh]•[Li]+
Cα77.476.4273.1878.12138.7
Cβ128.5128.82128.06130.34151.5
Sum (Cα + Cβ)205.9205.24201.24208.46290.2
Sum (Cβ − Cα)51.152.4073.1878.1212.8
PhPhPhPhPh
Ci150.6, 147.3150.33, 147.82151.51, 147.68152.59, 147.36159.6
Co123.3, 125.8132.47, 126.62132.88, 126.61132.75, 127.75136.4
Cm126.5, 132.4123.05, 125.98123.58, 126.61125.77, 125.93127.3
Cp107.8, 120.5107.65, 120.53108.49, 120.80110.87, 120.25124.3
Sum (Ci − Cp)/269.6/2 = 34.869.97/2=35.0069.90/2 = 34.9568.83/2 = 34.4235.3a
29Si-Ring-34.14-0.2713.69
CH3, tert-C2.5814.70, 18.88, 27.34, 29.0831.5, 33.3 (brd d)
Reference72 b73 bThis Work bThis Work b75 b
b In THF-d8, reference = 25.30 ppm.
Table 2. 13C/29Si-NMR chemical shifts of silole/germole dianions and silole anions.
Table 2. 13C/29Si-NMR chemical shifts of silole/germole dianions and silole anions.
[Ph4C4Si]2•2[Li]+ (2)[Ph4C4Ge]2•2[Li]+[Ph4C4Si( t-Bu)]•[Li]+[Ph4C4SiSiMe3]•[Li]+ (3)
Ring carbons151.22, 129.71 a165.57, 129.92 a155.76, 139.51159.67, 139.30
PhPh PhPh
Ci151.67, 145.83152.17, 146.30149.29, 144.72148.81, 145.72
Co129.97, 133.43129.92, 133.49130.50, 132.56129.81, 132.85
Cm126.38, 126.38126.38, 126.38126.40, 126.51126.30, 126.46
Cp119.48, 121.83119.29, 121.91121.38, 123.34120.86, 122.86
Sum(Ci − Cp)/256.19/2 = 28.1057.27/2 = 28.6449.29/2 = 24.6540.81/2 = 20.41
CH3, tert-C32.78(CH3), 23.58(tert-C)-0.23 [Si(CH3)3]
29Si-Ring68.5425.10−13.22
Refenence17 b18 b38 bThis Work b
a The reported assignments were revised [76], the chemical shifts did not coincided with each other [77]. b In THF-d8, reference = 25.30 ppm.
Upon adding trimethylsilyl chloride to 2 the chemical shifts of Cα and Cβ in 2 are shifted far downfield from 151.22 ppm and 129.71 ppm to 159.67 ppm and 139.30 ppm in 3. The chemical shifts of Ciα and Ciβ in 3 are observed at 145.72 ppm and 148.81 ppm, while the chemical shifts of C and C in 3 are observed at 122.86 ppm and 120.86 ppm respectively. These carbon peaks of four phenyl groups indicate that the phenyl groups of 3 are still polarized, and the average chemical shift difference of Ci and Cp is 20.41 ppm [Sum(Ci − Cp)/2] (Table 2). Such polarizations of phenyl groups are generally observed due to the absence of the significant π-conjugation between their phenyl groups and 5-membered ring because of their bulkiness and the congestion of four phenyl groups. The average chemical shift difference of 20.41 ppm for 3 is smaller than those of the silole dianion [Ph4C4Si]−2•2[Li]+ (2) (28.10 ppm) [17], [Ph4C4Si]−2•2[Na]+ (29.17 ppm) [16], the germole dianion [Ph4C4Ge]−2•2[Li]+ (28.64 ppm) [18], and even the silole anion [Ph4C4Si(t-Bu)]•[Li]+ (24.65 ppm) [38]. The difference is also significantly smaller than those of the localized 3-silolenes (the reduced siloles to 2,5-carbodianions); [Ph4C4SiMe2]−2•2[Li]+ (34.8 ppm) [72], [Ph4C4SiHMe]−2•2[Li]+ (35.00 ppm) [75], [Ph4C4Si(n-Bu)2]−2•2[Li]+ (34.95 ppm) (10), [Ph4C4Si(t-Bu)2]−2•2[Li]+ (34.42 ppm) (11), and that of the phenyl group in the localized germole anion [Me4C4GePh]•[Li]+ (35.3 ppm) [75] (Table 1). This trend implies that the electron density in the silole ring carbons of 3 is significantly lower than those in the rings of the localized 3-silolenes, the high aromatic silole/germole dianions {[Ph4C4Si]−2, [Ph4C4Ge]−2} and the silole anion [Ph4C4Si(t-Bu)] due to its low aromaticity and/or sp3 hydridization character on Si atom in 3.
X-ray crystallographic data for 1-trimethylsilyl-2,3,4,5-tetramethyl/ethyl-1-silacyclopentadienide anion) [R4C4Si(SiMe3)] (R = Me, Et) have revealed that the anionic rings possess a pyramidal silicon center and bond localization in the butadiene moiety of the ring, the 29Si-NMR chemical shifts of these pyramidal ring Si atoms in those anions are observed from −41 ppm to −54 ppm [23]. However in the case of 3 the 29Si-NMR chemical shift is observed at −13.22 ppm, far downfield from those of the pyramidal Si atoms in the localized silole anions and far upfield from those of silole dianions and silole anion (Table 3).
Table 3. 29Si-NMR chemical shifts of silole anions and dianion.
Table 3. 29Si-NMR chemical shifts of silole anions and dianion.
Silole Anion[Me4C4SiSiMe3]•[M]+[Et4C4SiSiMe3]• [M]+[Et4C4Si]2•2[M]+[Ph4C4SiSiMe3]•[M]+ (3)
MLiKLiKLiLi
29Si-Ring−45.38−43.96−42.70−41.52−53.12 c−47.3824.96−13.22
29Si-Ring with crown ether12-CE-418-CE-612-CE-4
29Si-SiMe3−12.47−11.68−12.44−11.00 −14.27−14.22−15.54
Reference22 a22 b22 a22 b22 c22 c51 cThis work c
a In CH2Cl2-d2. b In Benzene-d6. c In THF-d8.
The 13C-NMR and 29Si-NMR chemical shifts of 3 do not support its aromaticity, the introduction of trimethylsilyl group on the silicon atom might decrease aromaticity of silole anion with the substituent effect of the trimethylsilyl group enhancing the s-character of the lone pair on the silicon atom and decreasing the s-character of the Si-Si bond in 3 [23,48].

3. Experimental

General Procedures

All reactions were performed under an inert nitrogen atmosphere using standard Schlenk techniques. Air sensitive reagents were transferred in a nitrogen-filled glove box. THF and ether were distilled from sodium benzophenone ketyl under nitrogen. Hexane and pentane were stirred over concentrated H2SO4 and distilled from CaH2. NMR spectra were recorded on JEOL GSX270 and GSX400 spectrometers. GC-MS and solid sample MS data were obtained on a Hewlett-Packard 5988A GC-MS system equipped with a methyl silicon capillary column. Elemental analyses were done by Desert Analytics (Tucson, AZ, USA).
[Ph4C4Si(SiMe3)(R)] (R = Me (4), Et (5), i-Bu (6)). [Ph4C4SiCl2] (1) (0.57 g, 1.25 mmol) was sonicated in THF with an excess of lithium for 5 h. Then the remaining lithium was removed by filtration to give a red-purple solution. The solution was added to methyl iodide in THF with stirring at room temperature for 4 h to give a yellow solution. After removing the solvent under vacuum the remaining yellow solid was extracted with hexane. The concentrated solution was kept in a refrigerator for a couple of days to provide yellow crystals.
[Ph4C4Si(SiMe3)(Me)] (4). Yield: 0.38 g (65%); mp. 130–132 °C (lit. [40], mp. 130–132 °C).
[Ph4C4Si(SiMe3)(Et)] (5). Yield: 0.54 g (59%); mp. 100–102 °C, 1H-NMR (CDCl3, ref; ext. TMS = 0.00 ppm), 0.05 (s, SiMe3, 9H), 1.0–1.2 (brd m, ethyl, 5H), 6.68–7.15 (m, 20H); 13C-NMR (CDCl3, ref; solvent = 77.00 ppm), −1.40 (SiMe3), 3.06 (CH2), 8.64 (CH3); 29Si-NMR (CDCl3, ref; ext. TMS = 0.00), −2.53 (ring Si), −16.15 (SiMe3); Anal. Calcd. for C33H34Si2: C, 81.42; H, 7.04, Found: C, 81.59; H, 7.19.
[Ph4C4Si(SiMe3)(i-Bu)] (6). Yield: 0.44 g (68%); mp. 154–156 °C, 1H-NMR (CDCl3, ref; ext. TMS = 0.00 ppm), 0.03 (s, SiMe3, 9H), 0.92 (d, CMe2, 6H), 1.16 (d, CH2, 2H), 1.85 (m, CH, 1H), 6.68–7.15 (m, 20H); 29Si-NMR (CDCl3, ref; ext. TMS = 0.00), −6.49 (ring Si), −15.88 (SiMe3); Anal. Calcd. for C35H38Si2: C, 81.65; H, 7.44, Found: C, 81.76; H, 7.31.
[Ph4C4Si(n-Bu)2] (7). [Ph4C4SiCl2] (1) (0.57 g, 1.25 mmol) was sonicated with an excess of lithium for 5 h. Then the remaining lithium was removed by filtration to give a red-purple solution of the silole dianion. The solution was added to a THF solution of 1-bromobutane with stirring at room temperature for 10 h to give a yellow solution. After removing the solvent under vacuum the remaining yellow solid was extracted with hexane. The concentrated solution was kept in a refrigerator for a couple of days to provide yellow crystals. Yield: 0.56 g (90%); mp. 85 °C (lit. [78] mp. 81 °C).
[Ph4C4Si(t-Bu)2] (8). [Ph4C4SiCl2] (1) (0.55 g, 1.21 mmol) was sonicated with an excess of lithium in THF for 5 h. Then the remaining lithium was removed by filtration to give a red-purple solution of 2. The solution was added to a THF solution of t-butyl bromide with stirring at room temperature for 24 h to give a yellow solution. After removing the solvent under vacuum the remaining yellow solid was extracted with ether. The concentrated solution was kept in a refrigerator for a couple of days to provide pale yellow crystals of bissilole 1,1-bi[(t-Bu)SiC4Ph4] [14]. The filtered solution was concentrated under vacuum, then it was kept in a refrigerator for a couple of days to give yellow crystals of [Ph4C4Si(t-Bu)2] (8). Yield: 0.33 g (54%); mp. 169-171 °C, 1H-NMR (CDCl3, ref; ext. TMS = 0.00 ppm), 1.16 (s, Me, 18H), 6.68–7.15 (m, 20H), 29Si-NMR (CDCI3, ref; ext. TMS=0.00), 16.49 (ring Si); Anal. Calcd. for C36H38Si1: C,86.69; H,7,68, Found: C, 86.71; H, 7,75.
1,1-Bi[(t-Bu)SiC4Ph4] (9). Yield: 0.19 g (18%); mp. 295–307 °C (lit. [16] mp. 296-307 °C), 29Si-NMR (THF-d8, ref; ext. TMS = 0.00), 3.62 (ring Si).
1,1-Bis(R)-2,5-dilithio-2,3,4,5-tetraphenyl-1-silacyclopenta-3-enide anion [R = n-Bu (10), R = t-Bu (11)]. The respective [Ph4C4Si(n-Bu)2] (7) (0.025 g, 0.05 mmol) and [Ph4C4Si(t-Bu)2] (8) (0.025 g, 0.05 mmol) was transferred into 5 mm NMR tube, they were sonicated with an excess of lithium in THF-d8 for 2 h to give a red-purple solution. Then 1H-, 13C-, and 29Si-NMR spectroscopic study was performed.
[Ph4C4Si(n-Bu)2]-2•2[Li]+ (10); 1H-NMR (THF-d8, ref; ext. TMS = 0.00 ppm), 0.83 (t, CH3, 6H), 0.90 (m, CH2, 4H), 1.36 (sept, CH2, 4H), 1.52 (m, CH2, 4H), 6.68–7.15 (m, 20H), 29Si-NMR (THF-d8, ref; ext. TMS = 0.00), −0.27 (ring Si). [Ph4C4Si(t-Bu)2]-2•2[Li]+ (11); 1H-NMR (THF-d8, ref; ext. TMS = 0.00 ppm), 1.21 (brd s, Me, 18H), 6.68-7.15 (m, 20H), 29Si-NMR (THF-d8, ref; ext. TMS = 0.00), 13.69 (ring Si).

4. Conclusions

Silole dianion [Ph4C4Si]−2 (2) is a versatile intermediate to prepare symmetrically substituted siloles of [Ph4C4SiR2] (R = n-Bu, t-Bu) and unsymmetrically substituted siloles of [Ph4C4Si(SiMe3)(R)] (R = Me, Et, i-Bu). The formers are synthesized from the reaction of silole dianion 2 with the corresponding alkyl bromides, while the latter are synthesized via [Ph4C4Si(SiMe3)]•[Li] + (3) by hemsilylation of 2 with trimethylsilyl chloride and then by alkylation of 3 with the corresponding alkyl halides. The silole anion 3 and the reduced 3-silolenide 2,5-carbodianions {[Ph4C4Si(n-Bu)2]2•2[Li]+ (10) and [Ph4C4Si(t-Bu)2]2•2[Li]+ (11)} are characterized by 1H-, 13C-, and 29Si-NMR spectroscopy.

Acknowledgments

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NFR) funded by Ministry of Education. The author thanks Sung-Jun Cho (Pai Chai University, Daejon S. Korea) for supporting with lab facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Elschenbroich, C.; Salzer, A. Oranometallics-A Concise Introduction, 2nd ed.; VCH: Weinheim, Germany, 1992; pp. 315–343. [Google Scholar]
  2. Harder, S. Recent developments in cyclopentadienyl-alkalimetal chemistry. Cord. Chem. Rev. 1998, 176, 17–66. [Google Scholar] [CrossRef]
  3. Tongni, A.; Haterman, R. Metallocenes. Synthesis, Reactivity, Applications; VCH: Weinheim, Germany, 1998. [Google Scholar]
  4. Barton, T.J. Carbacyclic Silanes. In Comprehensive Organometallic Chemistry; Wilkinson, G., Stone, F.G.A., Abel, E.W., Eds.; Pergamon Press: Oxford, UK, 1982; pp. 205, 250–261. [Google Scholar]
  5. Dubac, J.; Laporterie, A.; Manuel, G. Group 14 metalloles. 1. Synthesis, organic chemistry, and physicochemical data. Chem. Rev. 1990, 90, 215–263. [Google Scholar] [CrossRef]
  6. Colomer, E.; Corriu, R.J.P.; Lheureux, M. Group 14 metalloles. 2. Ionic species and coordination compounds. Chem. Rev. 1990, 90, 265–282. [Google Scholar] [CrossRef]
  7. Dubac, J.; Guerin, C.; Meunier, P. The Chemistry of Organosilicon Compounds; Patai, S., Rappoport, Z., Eds.; Willey-Interscience: Chichester, UK, 1998; Volume 2, pp. 1009–1010. [Google Scholar]
  8. Gordon, M.S.; Boudjouk, P.; Anwari, F. Are the silacyclopentadienyl anion and the silacyclopropenyl cation aromatic? J. Am. Chem. Soc. 1983, 105, 4972–4976. [Google Scholar] [CrossRef]
  9. Damewood, J.R. Pyramidal inversion and electron delocalization in the silacyclopentadienyl anion. J. Org. Chem. 1986, 51, 5028–5029. [Google Scholar] [CrossRef]
  10. Goldfuss, B.; Schleyer, P.R. The Siloyl anion C4H4SiH is aromatic and the lithium silolide C4H4SiHLi even more so. Organometallics 1995, 14, 1553–1555. [Google Scholar] [CrossRef]
  11. Goldfuss, B.; Schleyer, P.R.; Hampel, F. Aromaticity in Silole Dianions: Structural, energetic, and magnetic aspects. Organometallics 1996, 15, 1755–1757. [Google Scholar] [CrossRef]
  12. Schleyer, P.R.; Maerker, C.; Dransfeld, A.; Jiao, H.; Hommes, N.J.R.V.E. Nucleus-independentchemical shifts: A simple and efficient aromaticity probe. J. Am. Chem. Soc. 1996, 118, 6317–6318. [Google Scholar]
  13. Goldfuss, B.; Schleyer, P.R. Aromaticity in group 14 metalloles: Structural, energetic, and magnetic criteria. Organometallics 1997, 16, 1543–1552. [Google Scholar] [CrossRef]
  14. Schleyer, P.R.; Manoharan, M.; Jiao, H.; Stahl, F. The Acenes: Is there a relationship between aromatic stabilization and reactivity? Org. Lett. 2001, 3, 3643–3646. [Google Scholar] [CrossRef]
  15. Wrackmeyer, B. Applications of 29Si NMR parameters. Annu. Rep. NMR Spectrosc. 2006, 57, 1–49. [Google Scholar] [CrossRef]
  16. Joo, W.-C.; Hong, J.-H.; Choi, S.-B.; Son, H.-E. Synthesis and reactivity of 1,1-disodio-2,3,4,5-tetraphenyl-1-silacyclopentadiene. J. Organomet. Chem. 1990, 391, 27–36. [Google Scholar] [CrossRef]
  17. Hong, J.-H.; Boudjouk, P.; Castellino, S. Synthesis and characterization of two aromatic siliconcontaining dianions: The 2,3,4,5-tetraphenylsilole dianion and the 1,1'-disila-2,2',3,3',4,4',5,5'-octaphenylfulvalene dianion. Organometallics 1994, 13, 3387–3389. [Google Scholar] [CrossRef]
  18. Hong, J.-H.; Boudjouk, P. Synthesis and characterization of a delocalized germanium-containing dianion: Dilithio-2,3,4,5-tetraphenyl-germole. Bull. Soc. Chim. Fr. 1995, 132, 495–498. [Google Scholar]
  19. West, R.; Sohn, H.; Bankwitz, U.; Calabrese, J.; Apeloig, T.; Mueller, T. Dilithium derivative of tetraphenylsilole: An η1-η5 dilithium structure. J. Am. Chem. Soc. 1995, 117, 11608–11609. [Google Scholar] [CrossRef]
  20. West, R.; Sohn, H.; Powell, D.R.; Mueller, T.; Apeloig, Y. Dianion of tetraphenylgermole is aromatic. Angew. Chem. Int. Ed. 1996, 35, 1002–1004. [Google Scholar] [CrossRef]
  21. Hong, J. H.; Pan, Y.; Boudjouk, P. A Novel Lithocenophane Derivative of a Trisgermole Dianion: [Li(thf)(tmeda)][2,3,4,5-Et4-Ge,Ge-{Li(2,3,4,5-Et4C4Ge)2}C4Ge]. Angew. Chem. Int. Ed. 1996, 35, 186–188. [Google Scholar] [CrossRef]
  22. Freeman, W.P.; Tilley, T.D.; Yap, G.P.A.; Rheingold, A.L. Siloyl anions and silole dianions: Structure of [K(18-crown-6)+]2[C4Me4Si2−]. Angew. Chem. Int. Ed. 1996, 35, 882–884. [Google Scholar] [CrossRef]
  23. Freeman, W.P.; Tilley, T.D.; Liable-Sands, L.M.; Rheingold, A.L. Synthesis and study of cyclic π-systems containing silicon and germanium. The question of aromaticity in cyclopentaidenyl analogues. J. Am. Chem. Soc. 1996, 118, 10457–10468. [Google Scholar] [CrossRef]
  24. Dysard, J.M.; Tilley, T.D. η5-Silolyl and η5-Germoyl complexes of d0 hafnium. Structural characterization of an η5-Silolyl complex. J. Am. Chem. Soc. 1998, 120, 8245–8246. [Google Scholar] [CrossRef]
  25. Choi, S.-B.; Boudjouk, P.; Hong, J.-H. Unique Bis-η5/η1bonding in a dianionic germole. synthesis and structural characterization of the dilithium salt of the 2,3,4,5-tetraethyl germole dianion. J. Am. Chem. Soc. 1999, 18, 2919–2921. [Google Scholar]
  26. Dysard, J.M.; Tilley, T.D. Hafnium-rhodium and hafnium-iridium heterobimetallic complexes featuring the bridging germole dianion ligand [GeC4Me4]2−. Organometallics 2000, 19, 2671–2675. [Google Scholar] [CrossRef]
  27. Dysard, J.M.; Tilley, T.D. Synthesis and reactivity of η5-Silolyl, η5-Germoyl, and η5-Germoyl dianion complexes of zirconium and hafnium. J. Am. Chem. Soc. 2000, 122, 3097–3105. [Google Scholar] [CrossRef]
  28. Saito, M.; Haga, M.; Yoshioka, M. Formation of the first monoanion and dianion of stannole. Chem. Commun. 2002, 1002–1003. [Google Scholar] [CrossRef]
  29. Saito, M.; Haga, M.; Yoshioka, M.; Ishimura, K.; Nagase, S. The Aromaticity of the stannole dianion. Angew. Chem. Int. Ed. 2005, 44, 6553–6556. [Google Scholar] [CrossRef]
  30. Saito, M.; Kuwabara, T.; Kambayashi, C.; Yoshioka, M.; Ishimura, K.; Nagase, S. Synthesis, structure, and reaction of tetraethyldilithiostannole. Chem. Lett. 2010, 39, 700–701. [Google Scholar] [CrossRef]
  31. Saito, M.; Sakaguchi, M.; Tajima, T.; Ishimura, K.; Nagase, S.; Hada, M. Dilithioplumbole: A lead-bearing aromatic cyclopentadienyl analog. Science 2010, 328, 339–342. [Google Scholar] [CrossRef]
  32. Lee, V.Y.; Sekiguchi, A.; Ichinohe, M.; Fukaya, N. Stable aromatic compounds containing heavier group 14 elements. J. Organomet. Chem. 2000, 611, 228–235. [Google Scholar] [CrossRef]
  33. Hissler, M.; Dyer, P.W.; Reau, R. Linear organic π-conjugated systems featuring the heavy group 14 and 15 elements. Coord. Chem. Rev. 2003, 244, 1–44. [Google Scholar] [CrossRef]
  34. Grützmacher, H. Five-membered aromatic and antianromatic rings with gallium, germanium, and bismuth centers: A short march through the periodic table. Angw. Chem. Int. Ed. 2005, 34, 295–298. [Google Scholar]
  35. Saito, M.; Yoshioka, M. The anions and dianions of group 14 metalloles. Coord. Chem. Rev. 2005, 249, 765–780. [Google Scholar] [CrossRef]
  36. Lee, V.Y.; Sekiguchi, A. Aromaticity of group 14 organometallics: Experimental aspects. Angw. Chem. Int. Ed. 2007, 46, 6596–6620. [Google Scholar] [CrossRef]
  37. Saito, M. Challenge to expand the concept of aromaticity to tin- and lead-containing carbocyclic compounds: Synthesis, structures and reactions of dilithiostannoles and dilithioplumbole. Coord. Chem. Rev. 2012, 256, 627–636. [Google Scholar] [CrossRef]
  38. Hong, J.-H.; Boudjouk, P. A stable aromatic species containing silicon. Synthesis and characterization of the 1-tert-Butyl-2,3,4,5-tetraphenyl-1-silacyclopentadienide anion. J. Am. Chem. Soc. 1993, 115, 5883–5884. [Google Scholar] [CrossRef]
  39. Sohn, H.; Powel, R.; West, R.; Hong, J.-H.; Joo, W.-C. Dimerization of the silole anion [C4Ph4SiMe] to a tricyclic diallylic dianion. Organometallics 1997, 16, 2770–2772. [Google Scholar] [CrossRef]
  40. Hong, J.-H. Dissociation of the disilatricyclic diallylic dianion [(C4Ph4SiMe)2]−2 to the silole anion [MeSiC4Ph4] by halide ion coordination or halide ion nucleophilic substitution at the silicon atom. Molecules 2011, 16, 8451–8462. [Google Scholar] [CrossRef]
  41. Lee, V.Y.; Kato, R.; Ichinohe, M.; Sekiguchi, A. The heavy analogue of CpLi: Lithium 1,2-disila-3-germacyclopentadienide, a 6π-electron aromatic system. J. Am. Chem. Soc. 2005, 127, 13142–13143. [Google Scholar] [CrossRef]
  42. Yasuda, H.; Lee, V.Y.; Sekiguchi, A. Si3C2-Rings: From a nonconjugated trisilacyclopentadiene to an aromatic trisilacyclopentadienide and cyclic disilenide. J. Am. Chem. Soc. 2009, 131, 6352–6353. [Google Scholar] [CrossRef]
  43. Wang, W.; Yao, S.; Wüllen, C.V.; Driess, M. A cyclopentadienide analogues containing divalent germanium and a heavy cyclobutadiene-like dianion with an unusual Ge4 core. J. Am. Chem. Soc. 2008, 130, 9640–9641. [Google Scholar] [CrossRef]
  44. Lee, V.Y.; Kato, R.; Sekiguchi, A.; Krapp, A.; Frenking, G. Heavy ferrocene: A sandwich complex containing Si and Ge atoms. J. Am. Chem. Soc. 2007, 129, 10340–10341. [Google Scholar] [CrossRef]
  45. Yasuda, H.; Lee, V.Y.; Sekiguchi, A. η5-1,2,3-Trisilacyclopentadienyl—A ligand for transition metal complexes: Rhodium half-sandwich and ruthenium sandwich. J. Am. Chem. Soc. 2009, 131, 9902–9903. [Google Scholar] [CrossRef]
  46. Freeman, W.P.; Tilley, T.D. Stable Silacyclopentadienyl complexes of ruthenium: (η5-C5Me5)Ru[η5-C5Me5SiSi(SiMe3)3] and X-ray structure of its protonated form. J. Am. Chem. Soc. 1994, 116, 8428–8429. [Google Scholar] [CrossRef]
  47. Freeman, W.P.; Dysard, J.M.; Tilley, T.D.; Rheingold, A.L. Synthesis and reactivity of η5-germacyclopentadienyl complexes of iron. Organometallics 2002, 21, 1734–1738. [Google Scholar] [CrossRef]
  48. Saito, M.; kuwabara, T.; Ishimura, K.; Nagase, S. Synthesis and structures of lithium salts of stannole anions. Bull. Chem. Soc. Jpn. 2010, 83, 825–827. [Google Scholar] [CrossRef]
  49. Braye, E.H.; Hübel, W.; Caplier, I. New unsaturated heterocyclic systems. I. J. Am. Chem. Soc. 1961, 83, 4406–4413. [Google Scholar] [CrossRef]
  50. Bankwitz, U.; Sohn, H.; Powell, D.R.; West, R. Synthesis, soilid-state structure, and reduction of 1,1-dichloro-2,3,4,5-tetramethylsilole. J. Organomet. Chem. 1995, 499, C7–C9. [Google Scholar] [CrossRef]
  51. Hong, J.H. Synthesis and NMR-study of the 2,3,4,5-tetraethylsilole dianion [SiC4Et4]2−•2[Li]+. Molecules 2011, 16, 8033–8040. [Google Scholar] [CrossRef]
  52. Tamao, K.; Yamaguchi, S.; Shiro, M. Oligosiloles: First synthesis based on a novel endo-endo modeintramolecular reductive cyclization of diethynylsilanes. J. Am. Chem. Soc. 1994, 116, 11715–11722. [Google Scholar] [CrossRef]
  53. Yamaguchi, S.; Jin, R.-Z.; Tamao, K.; Shiro, M. Synthesis of a series of 1,1-difunctionalized siloles. Organometallics 1997, 16, 2230–2232. [Google Scholar] [CrossRef]
  54. Fagan, P.J.; Nugent, W.A.; Calabrese, J.C. Metallacycle transfer from zirconium to main group element: A versatile synthesis of heterocycles. J. Am. Chem. Soc. 1994, 116, 1880–1889. [Google Scholar] [CrossRef]
  55. Wrackmeyer, B. Metallacyclopentadienes and related heterocycles via 1,1-organoboration of alkyn-1-ylmetal compounds. Heteroat. Chem. 2006, 17, 188–208. [Google Scholar] [CrossRef]
  56. Wrackmeyer, B.; Tok, O.L. Comprehensive Heterocyclic Chemistry III; Katritzky, A.R., Ramsden, C.A., Scriven, E.F.V., Taylor, R.J.K., Eds.; Elsevier: Oxford, UK, 2008; Chapter 3.17; pp. 1181–1223. [Google Scholar]
  57. Khan, E.; Bayer, S.; Kempe, R.; Wrackmeyer, B. Synthesis and molecular structure of silole derivatives bearing functional groups on silicon: 1,1-organoboration of dialkynylsilanes. Eur. J. Inorg. Chem. 2009, 4416–4424. [Google Scholar]
  58. Dierker, G.; Ugolotti, J.; Kehr, G.; Fröhlich, R.; Erker, G. Reaction of bis(alkynyl)silanes with tris(pentafluorophenyl)borane: Synthesis of bulky silole derivatives by means of 1,1-carboboration under mild reaction conditions. Adv. Synth. Catal. 2009, 351, 1080–1088. [Google Scholar] [CrossRef]
  59. Ishigawa, M.; Tabohashi, T.; Sugisawa, H.; Nishimura, K.; Kumada, M. Chemistry of siloles. the reaction of siloles with organolithium reagents. J. Orgamomet. Chem. 1983, 205, 109–119. [Google Scholar]
  60. Ishigawa, M.; Tabohashi, T.; Ohashi, H.; Kumada, M.; Iyoda, J. Chemistry of siloles. 1-methyldibenzosilacyclopentadienide anion. Orgamometallics 1983, 2, 351–352. [Google Scholar] [CrossRef]
  61. Yamaguchi, S.; Jin, R.-Z.; Tamao, K. Silicon-catenated silole oligomers: Oligo(1,1-silole)s. Organometallics 1997, 16, 2486–2488. [Google Scholar] [CrossRef]
  62. Sohn, H.; Merritt, J.; Powell, D.R.; West, R. A new spirocyclic system: Synthesis of a silaspirotropylidene. Organometallics 1997, 16, 5133–5134. [Google Scholar] [CrossRef]
  63. Sanji, T.; Sakai, T.; Kabuto, C.; Sakurai, H. Silole-incorporated polysilanes. J. Am. Chem. Soc. 1998, 120, 4552–4553. [Google Scholar] [CrossRef]
  64. Kanno, K.; Ichinohe, M.; Kabuto, C.; Kira, M. Synthesis and structure of a series of oligo [1,1-(2,3,4,5-tetramethylsilole)]s. Chem. Lett. 1998, 27, 99–100. [Google Scholar]
  65. Yamaguchi, S.; Jin, R.-Z.; Tamao, K. Silole polymer and cyclic hexamer catenating through the ring silicones. J. Am. Chem. Soc. 1999, 121, 2937–3938. [Google Scholar] [CrossRef]
  66. Yamaguchi, S.; Jin, R.-Z.; Tamao, K. New type of polysilanes: Poly(1,1-silole)s. J. Organomet. Chem. 2000, 611, 5–11. [Google Scholar]
  67. Saito, M.; Haga, M.; Yoshioka, M. Synthesis of stannole anion by alkylation of stannole dianion. Chem. Lett. 2003, 32, 912–913. [Google Scholar] [CrossRef]
  68. Haga, R.; Saito, M.; Yoshioka, M. Synthesis and reactions of stannole anions. Eur. J. Inorg. Chem. 2007, 1297–1306. [Google Scholar] [CrossRef]
  69. Haga, R.; Saito, M.; Yoshioka, M. Reversible redox behavior between stannole dianion and bistannole-1,2-dianion. J. Am. Chem. Soc. 2006, 128, 4934–4935. [Google Scholar] [CrossRef]
  70. Haga, R.; Saito, M.; Yoshioka, M. Stepwise oxidation of the stannole dianion. Chem. Eur. J. 2008, 14, 4068–4073. [Google Scholar] [CrossRef]
  71. Schäfer, A.; Weidenbruch, M.; Pohl, S. Siliciumverbindungen mit starken intramolekularen sterischen wechselwirkungen: XIX. Simultane bildung und reaktionen von di-t-butylsilandiyl und tetra-t-butyldisilen. J. Organomet. Chem. 1985, 282, 305–313. [Google Scholar] [CrossRef]
  72. O’Brien, D.H.; Breeden, D.L. Tetraanion of 1,1-dimethyl-2,3,4,5-tetraphenyl-1-silacyclopentadiene. J. Am. Chem. Soc. 1981, 103, 3237–3239. [Google Scholar] [CrossRef]
  73. Wakahara, T.; Ando, W. Reaction of hydrosilanes with lithium. Formation of silole anions from 1-methylsilole via carbodianion. Chem. Lett. 1997, 11, 1179–1180. [Google Scholar] [CrossRef]
  74. Hong, J.-H.; Boudjouk, P. Synthesis and characterization of a novel pentavalent silane: 1-methyl-1,1-dihydrido-2,3,4,5-tetraphenyl-1-silacyclopentadiene silicate, [Ph4C4SiMeH2][K+]. Organometallics 1995, 14, 574–576. [Google Scholar] [CrossRef]
  75. Dufour, P.; Dubac, J.; Dartiguenave, M.; Dartiguenave, Y. C-methylated (germacyclopentadienyl) lithium. Organometallics 1990, 9, 3001–3003. [Google Scholar] [CrossRef]
  76. Tandura, S.N.; Troitskii, N.A.; Kolesnikov, S.P.; Nosov, K.S.; Egorov, M.P. 13C NMR study of 2,3,4,5-tetraphenyl silole dilithium salt. Russ. Chem. Bull. 1999, 48, 214–217. [Google Scholar] [CrossRef]
  77. Tandura, S.N.; Kolesnikov, S.P.; Nosov, K.S.; Egorov, M.P.; Nefedov, O.M. Charge localization in the dianion of tetraphenylgermole according to 13C NMR data. Main Group Met. Chem. 1999, 22, 9–14. [Google Scholar]
  78. Jutzi, P.; Karl, A. Synthese und reaktionen einiger 1R,1R’-2,3,4,5-tetraphenyl-1-silacyclopentadiene. J. Organomet. Chem. 1981, 214, 289–302. [Google Scholar] [CrossRef]
  • Sample Availability: Not Available.

Share and Cite

MDPI and ACS Style

Hong, J.-H. Synthesis and NMR-Study of 1-Trimethylsilyl Substituted Silole Anion [Ph4C4Si(SiMe3)]•[Li]+ and 3-Silolenide 2,5-carbodianions {[Ph4C4Si(n-Bu)2]2•2[Li]+, [Ph4C4Si(t-Bu)2]2•2[Li]+} via Silole Dianion [Ph4C4Si]2•2[Li]+. Molecules 2013, 18, 10568-10579. https://doi.org/10.3390/molecules180910568

AMA Style

Hong J-H. Synthesis and NMR-Study of 1-Trimethylsilyl Substituted Silole Anion [Ph4C4Si(SiMe3)]•[Li]+ and 3-Silolenide 2,5-carbodianions {[Ph4C4Si(n-Bu)2]2•2[Li]+, [Ph4C4Si(t-Bu)2]2•2[Li]+} via Silole Dianion [Ph4C4Si]2•2[Li]+. Molecules. 2013; 18(9):10568-10579. https://doi.org/10.3390/molecules180910568

Chicago/Turabian Style

Hong, Jang-Hwan. 2013. "Synthesis and NMR-Study of 1-Trimethylsilyl Substituted Silole Anion [Ph4C4Si(SiMe3)]•[Li]+ and 3-Silolenide 2,5-carbodianions {[Ph4C4Si(n-Bu)2]2•2[Li]+, [Ph4C4Si(t-Bu)2]2•2[Li]+} via Silole Dianion [Ph4C4Si]2•2[Li]+" Molecules 18, no. 9: 10568-10579. https://doi.org/10.3390/molecules180910568

Article Metrics

Back to TopTop