首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
An experimental and theoretical study of the base‐stabilized disilene 1 is reported, which forms at low temperatures in the disproportionation reaction of Si2Cl6 or neo‐Si5Cl12 with equimolar amounts of NMe2Et. Single‐crystal X‐ray diffraction and quantum‐chemical bonding analysis disclose an unprecedented structure in silicon chemistry featuring a dative Si→Si single bond between two silylene moieties, Me2EtN→SiCl2→Si(SiCl3)2. The central ambiphilic SiCl2 group is linked by dative bonds to the amine donor and the bis(trichlorosilyl)silylene acceptor, which leads to push–pull stabilization. Based on experimental and theoretical examinations a formation mechanism is presented that involves an autocatalytic reaction of the intermediately formed anion Si(SiCl3)3? with neo‐Si5Cl12 to yield 1 .  相似文献   

2.
Closely following the procedure for the preparation of the base‐stabilized dichlorosilylene complex NHCDipp⋅SiCl2 reported by Roesky, Stalke, and co‐workers (Angew. Chem. Int. Ed . 2009 , 48 , 5683–5686), a few crystals of the salt [NHCDipp−H⋅⋅⋅Cl⋅⋅⋅H−NHCDipp]Si(SiCl3)3 were isolated, aside from the reported byproduct [NHCDipp−H+⋅⋅⋅Cl], and characterized by X‐ray crystallography (NHCDipp=N,N‐di(2,6‐diisopropylphenyl)imidazo‐2‐ylidene). They contain the weakly coordinating anion Si(SiCl3)3, which was also obtained in high yields upon deprotonation of the conjugate Brønsted acid HSi(SiCl3)3 with NHCDipp or PMP (PMP=1,2,2,6,6‐pentamethylpiperidine). The acidity of HSi(SiCl3)3 was estimated by DFT calculations to be substantially higher than those of other H‐silanes. Further DFT studies on the electronic structure of Si(SiCl3)3, including the electrostatic potential and the electron localizability, confirmed its low basicity and nucleophilicity compared with other silyl anions.  相似文献   

3.
Tetrakis(trichlorosilyl)silan (neo-Si5Cl12) is cleaved in an amine catalysed reaction by HCl in benzene solution to tris(trichlorosilyl)silan HSi(SiCl3)3 (1). The amine catalysed cleavage of1 with different amines and solvents is investigated. A new method for preparation of pentachlorodisilane HSi2Cl5 is described and a reaction mechanism for the cleavage is postulated.
  相似文献   

4.
Infrared andRaman vibrational spectra ofn-Si4Cl10,n-Si5Cl12,neo-Si5Cl12 and [(SiCl3)3Si]2 have been measured and assigned. A local symmetry force field has been developed to simulate vibrational spectra of all (noncyclic) perchlorosilanes Si n Cl2n+2 known today (n=2, 3, 4, 5, 8). The observed spectra are reproduced satisfactorily
Die Vibrationsspektren linearer und verzweigter PerchlorsilaneSi n Cl 2n+2 und deren Simulierung mittels eines lokalen Symmetrie-Kraftfeldes
Zusammenfassung Infrarot- undRaman-Spektren vonn-Si4Cl10,n-Si5Cl12,neo-Si5Cl12 und [(SiCl3)3Si]2 wurden aufgenommen und zugeordnet. Ein lokales Symmetrie-Kraftfeld zur Simulation der Spektren aller bisher bekannten (nicht cyclischen) Perchlorsilane Si n Cl2n+2 (n=2, 3, 4, 5, 8) wird angegeben. Die beobachteten Spektren werden zufriedenstellend reproduziert
  相似文献   

5.
The reaction of a metastable SiCl2 solution with the sterically less‐demanding carbene N,N‐diisopropylimidazo‐2‐ylidene (IPr) yields the salt [(IPr3Si3Cl5)+]Cl? ( 1 ‐Cl), containing a silyl cation with a Si3 backbone. Salt 1 is highly reactive, but it can be used as a reagent in deuterated dichloromethane, whereby dehalogenation with Me3SiOTf (OTf=O3SCF3) gives the dicationic silyl halide [(IPr3Si3Cl4)]2+ 2 . Quantum chemical calculations show that the HOMO is localized at the negatively charged central silicon atom of 1 and 2 , and thus although both compounds are cations they are better described as silanides, which was also corroborated by NMR investigations.  相似文献   

6.
Incorporation of SiIV into an expanded porphyrin has been achieved for the first time. Treatment of [28]hexaphyrin 1 with CH3SiCl3 and N,N‐diisopropylethylamine gave SiIV complex 2 and its N‐fused product 4 that both have Möbius aromatic nature. In both complexes, the coordinated Si atom is satisfied in a typical trigonal bipyramidal coordination. SiIV incorporation induces conformational rigidification and redshifted absorption profiles due to σ–π conjugation between the Si atom and hexaphyrin macrocycle. Tamao–Fleming oxidation of 2 with H2O2 gave β‐hydroxy [28]hexaphyrin 5 , which exists as a ruffled rectangular shape in the solid state, yet it has been revealed to exist predominantly as a twisted Möbius aromatic conformer in CH2Cl2.  相似文献   

7.
Schröder  D.  Schwarz  H. 《Russian Chemical Bulletin》2001,50(11):2087-2091
Sector-field mass spectrometry was used to probe the fragmentation patterns of the cationic silicon chlorides Si2Cln + (n = 1—6). For almost all Si2Cln + ions, Si—Si fragmentation predominates the Si—Cl bond cleavage both in the metastable ion and collisional activation mass spectra. Analysis of the fragmentation patterns indicates that the long-lived radical cation Si2Cl6 ·+ corresponds to a complex [SiCl2·SiCl4]·+ rather than the intact molecular ion of hexachlorodisilane. The behavior of Si2Cl5 + is consistent with the formation of the (trichlorosilyl)dichlorosilyl cation Cl3SiSICl2 +. Structural aspects are also discussed for the other Si2Cln + species. A semi-quantitative analysis of the fragmentation patterns in conjunction with the literature thermochemistry data was used to estimate some thermochemical properties of the Si2Cln + cations.  相似文献   

8.
Formation of Organosilicon Compounds. 111. The Hydrogenation of Si-chlorinated, C-spiro-linked 2,4-Disilacyclobutanes with LiAlH4 or iBu2AlH. The Access to Si8C3H20 The hydrogenation of Si-chlorinated, C-spiro-linked 2,4-disilacyclobutanes containing C(SiCl3)2 terminal groups with LiAlH4 in Et2O proceeds under complete cleavage of the fourmembered rings and under elimination of one SiH3 group. Such, Si8C3Cl20 4 forms (H3Si)2CH? SiH2? CH(SiH3)? SiH2? CH(SiH3)2 4 α, and even Si8C3H20 4a with LiAlH4 forms 4 α. The hydrogenation of related compounds containing however CH(SiCl3) terminal groups similarly proceeds under ring cleavage but no SiH3 groups are eliminated. Such, (Cl3Si)CH(SiCl2)2CH(SiCl3) 41 forms (H3Si)2CH? SiH2? CH2(SiH3) 41 α. However, in reactions with iBu2AlH in pentane neither the disilacyclobutane rings are cleaved nor are SiH3 groups eliminated. Only by this method Si8C3H20 is accessible from 4 , Si6C2H16 3a from Si6C2Cl16 3 and Si4C2H12 41a from 41 . C(SiCl3)4 cleanly produces C(SiH3)4. Based on the knowledge about the different properties of LiAlH4 and iBu2AlH in hydrogenation reactions of disilacyclo-butanes it was possible to elucidate the composition and the structures of the hydrogenated derivatives of the product mixture from the reaction of MeCl2Si? CCl2? SiCl3 with Si(Cu) [1] and to trace them back to the initially formed Si chlorinated disilacyclobutanes Si6C2Cl15Me 34 , Si6C2Cl14Me2 35 , Si8C3Cl19Me 36 and Si8C3Cl18Me2 37 . Compound 4a forms colourless crystals of space group P1 with a = 799.7(6), b = 1263.6(12), c = 1758.7(14) pm, α = 103.33(7)°, β = 95.28(6)°, γ = 105.57(7)° and Z = 4.  相似文献   

9.
The reaction of 1,3‐diisopropylimidazolin‐2‐ylidene (iPr2Im) with diphenyldichlorosilane (Ph2SiCl2) leads to the adduct (iPr2Im)SiCl2Ph2 1 . Prolonged heating of isolated 1 at 66 °C in THF affords the backbone‐tethered bis(imidazolium) salt [(aHiPr2Im)2SiPh2]2+ 2 Cl? 2 (“a” denotes “abnormal” coordination of the NHC), which can be synthesized in high yields in one step starting from two equivalents of iPr2Im and Ph2SiCl2. Imidazolium salt 2 can be deprotonated in THF at room temperature using sodium hydride as a base and catalytic amounts of sodium tert‐butoxide to give the stable N‐heterocyclic dicarbene (aiPr2Im)2SiPh2 3 , in which two NHCs are backbone‐tethered with a SiPh2 group. This easy‐to‐synthesize dicarbene 3 can be used as a novel ligand type in transition metal chemistry for the preparation of dinuclear NHC complexes, as exemplified by the synthesis of the homodinuclear copper(I) complex [{a(ClCu?iPr2Im)}2SiPh2] 4 .  相似文献   

10.
Formation of Organosilicon Compounds. 110. Reactions of (Cl3Si)2CCl2 and its Si-methylated Derivatives as well as of (Cl3Si)2CHCl, (Cl3Si)2C(Cl)Me and Me2CCl2 with Silicon (Cu cat.) The reactions of (Cl3Si)2CCl2 1 , its Si-methylated derivatives (Me3Si)2CCl2 8 , Me3Si? CCl2? SiMe2Cl 9 , (ClMe2Si)2CCl2 10 , Me3Si? CCl2? SiMeCl2 11 , Cl2MeSi? CCl2? SiCl3 12 as well as of (Cl3Si)2CHCl 38 , (Cl3Si)2CClMe 39 and of Me2CCl2 with Si (Cu cat.) in a fluid bed reactor ( 38 and 39 also in a stirred solid bedreactor) arc presented. While (Cl3Si)2CCl2 1 yields C(SiCl3)4 2 the 1,1,3,3-tetrachloro-2,2,4,4-tetrakis(trichlorsilyl)-1,3-disilacyclobutane Si6C2Cl16 3 and the related C-spiro linked disilacyclobutanes Si8C3Cl20 4 , Si10C4Cl24 5 , Si12C5Cl28 6 , Si14C6Cl32 7 this type of compounds is not obtained starting from the Si-methylated derivatives 8, 9, 10, 11 They Produce a number of variously Si-chlorinated and -methylated tetrasila- and trisilamethanes. However, Cl2MeSi? CCl2? SiCl3 12 forms besides of Si-chlorinated trisilamethanes also the disilacyclobutanes Si6C2Cl15Me 34 and cis- and trans Si6C2Cl14Me2 35 as well as the spiro-linked disilacyclobutanes Si8C3Cl19Me 36 , Si8C3Cl18Me2 37 . (Cl3Si)2CHCl 38 mainly yields HC(SiCl3)3 31 and also the disilacyclobutanes cis- and trans-(Cl3Si)HC(SiCl2)2CH(SiCl3) 41 and (Cl3Si)2C(SiCl2)2CH(SiCl3) 45 the 1,3,5-trisilacyclohexane [Cl3Si(H)C? SiCl2]3 44 as well as [(Cl3Si)2CH]2SiCl2, and (Cl3Si)2CClMe 39 mainly yields (Cl3Si)2C?CH2and (Cl3Si)2besides of HC(SiCl3)3, MeC(SiCl3)3and (Cl3Si)3C? SiCl2Me.,. Me2CCl2 59 mainly yields Me(Cl)C?CH2, Me2CHCl and HCl2Si? CMe2? SiCl3, besides of Me2C(SiCl3)2 and Me2C(SiCl2H)2 Compound 3 crystallizes triclinically in the space group P1 (Nr. 2) mit a = 900,3, b = 914,0, c = 855,3 pm, α = 116,45°, β = 101,44°, γ = 95,86° and one molecule per unit cell. Compound 4 crystallizes monoclinically in thc space group C2/c (no. 15) with a = 3158.3,b = I 103.7, c = 2037.4 pm, β = 1 16.62° and 8 molecules pcr unit cell. The disilacyclobutane ring of compound 3 is plane, showing a mean distance of d (Si-C) =19 1.8 pm and the usual deformations of endocyclic angles: αSi = 94,2°> 85,8° = αC.The spiro-linked disilacyclobutane rings of compound 4 are slightly folded by a mean angle of (19.0°). Their mean distances were found to be d (Si? C) = 190.4 pm relating to the central carbon atom and 192.0 pm to the outer ones, respectively. The deformations of endocyclic angles: αSi = 93,9°> 84,4° = αC are comparable to those of compound 3.  相似文献   

11.
Thermal silazane cleavage of dichloroboryldisilylamines (SiClmMe3?m)N(SiMe3)(BCl2) (1: m = 1; 2: m = 2) at 196 °C leads to the borazine derivates [(SiClmMe3?m)NB(ClnMe1?n)]3 (3: m = 1, n = 0.185; 4: m = 2, n = 0.111) characterized by NMR and IR spectroscopy and mass spectrometry. Single‐crystal X‐ray diffraction structure analyses reveal (BN)3 units with unusual twisted boat conformations in both compounds. Additionally, more detailed studies are done to clear up the function of the by‐products (SiClmMe3?m)N(SiClMe2)(BClMe) formed during the cyclization step leading to asymmetrically boron substituted borazine derivatives. The single‐source precursors 3 and 4 were cross‐linked with methylamine producing polymers 3P and 4P, which were transformed into black amorphous materials with ceramic yields of 20.8 % and 50.3 %, respectively. Ceramic 4C (Si1.00B0.98 N2.55 C1.37O0.05) was further investigated by 11B and 29Si magic angle spinning (MAS) NMR spectroscopy. A combined study of high‐temperature TG analyses and X‐ray powder diffraction analyses confirms the thermal stability of 4C up to 1670 °C. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

12.
Contributions to the Chemistry of Halogenosilane Adducts. VII. 2,2′-Bipyridine-mono-N-oxide Complexes of Halogenosilanes The new adducts SiF4 · bipyO, SiCl4 · bipyO, SiCl4 · 2 bipyO, SiBr4 · 2 bipyO, SiI4 · 2 bipyO and SiI4 · 3 bipyO have been prepared by the reaction of SiF4, SiCl4, SiBr4 and SiI4 with bipyO in the appropriate molar ratio. The 1:1 reaction of Si2Cl6 and bipyO yields the adduct Cl3SiOSiCl3 · bipy (reduction of the amineoxide), in a 1:2 molar ratio Cl3SiOSiCl3 · bipy and Cl3SiOSiCl3 · 2 bipyO are obtained. The latter compound is formed in the reaction of Cl3SiOSiCl3 · bipy and bipyO and also in the reaction of Cl3SiOSiCl3 and bipyO. The results show that bipyO is a stronger base than bipy with reference to the silanes investigated. Some properties of the compounds and structural investigations are reported. bipyO is chelating and silicon is hexacoordinated in all adducts. The following structures are suggested: SiX4 · bipyO (molecular), [SiX2 · 2 bipyO]X2 (ionic), [Si · 3 bipyO]I4 (ionic). In Cl3SiOSiCl3 · 2 bipyO both ligands are coordinated to one Si atom.  相似文献   

13.
In the adduct ferrocene‐1,1′‐diyl­bis­(di­phenyl­methanol)–1,2‐bis(4‐pyridyl)­ethene (1/1), [Fe(C18H15O)2]·C12H10N2, there is an intramolecular O—H?O hydrogen bond in the ferro­cene­diol component and a single O—H?N hydrogen bond linking the diol to the di­amine, which is disordered over two sets of sites, so forming a finite monomeric adduct. In the adduct ferrocene‐1,1′‐diyl­bis­(di­phenyl­methanol)–1,6‐di­amino­hexane (2/1), 2[Fe(C18H15O)2]·C6H16N2, the amine lies across a centre of inversion in space group P. There is an intramolecular O—H?O hydrogen bond in the ferrocenediol, and the molecular components are linked by O—H?N and N—H?O hydrogen bonds, one of each type, into a C(13)[R(12)] chain of rings.  相似文献   

14.
The crystal structure of the title compound, C19H26NO+·Cl? (common name: N,N‐diethyl‐2‐[(4‐phenyl­methyl)phenoxy]‐ethan­amine hydro­chloride), contains one mol­ecule in the asymmetric unit. The planes through the two phenyl rings are roughly perpendicular. Protonation occurs at the N atom, to which the Cl? ion is linked via an N—H?Cl hydrogen bond. The mol­ecule adopts an eclipsed rather than extended conformation.  相似文献   

15.
The X‐ray structure determinations of the two title com­pounds, namely 7‐methyl‐7,17‐di­aza‐3,11‐diazo­niabi­cyclo[11.3.1]­hep­ta­deca‐1(17),13,15‐triene dichloride monohydrate, C14H26N42+·2Cl?·H2O, (I), and 7‐methyl‐17‐aza‐3,7,11‐triazo­niabi­cyclo­[11.3.1]­heptadeca‐1(17),13,15‐triene 2.826‐chloride 0.174‐nitrate, C14H27N43+·2.826Cl?·0.174NO3?, (II), are re­ported. Protonation occurs at the secondary amine N atoms in (I) and at all three amine N atoms in (II) to which the Cl? ions are linked via N—H?Cl hydrogen bonds. The macrocyclic hole is quite different in both structures, as is observed by comparing particularly the N3?N4 distances [2.976 (4) and 4.175 (4) Å for (I) and (II), respectively]. In (II), a Cl? ion alternates with an NO3? ion in a disordered structure.  相似文献   

16.
A novel tetraoxolene‐bridged Fe two‐dimensional honeycomb layered compound, (NPr4)2[Fe2(Cl2An)3] ?2 (acetone)?H2O ( 1 ), where Cl2Ann?=2,5‐dichloro‐3,6‐dihydroxy‐1,4‐benzoquinonate and NPr4+=tetrapropylammonium cation, has been synthesized. 1 revealed a thermally induced valence tautomeric transition at T1/2=236 K (cooling)/237 K (heating) between Fem+ (m=2 or 3) and Cl2Ann? (n=2 or 3) that induced valence modulations between [FeIIHSFeIIIHS(Cl2An2?)2(Cl2An.3?)]2? at T>T1/2 and [FeIIIHSFeIIIHS(Cl2An2?)(Cl2An.3?)2]2? at T<T1/2. Even in a two‐dimensional network structure, the low‐temperature phase [FeIIIHSFeIIIHS(Cl2An2?)(Cl2An.3?)2]2? valence set can be regarded as a magnetic chain‐knit network, where ferrimagnetic Δ and Λ chains of [FeIIIHS(Cl2An.3?)] are alternately linked by the diamagnetic Cl2An2?. This results in a slow magnetization behavior attributed to the structure acting as a single‐chain magnet at lower temperatures.  相似文献   

17.
Silicon analogues of the most prominent carbon nanostructures, namely, hollow spheroidals such as C60 and the fullerene family, have been unknown to date. Herein we show that discrete Si20 dodecahedra, stabilized by an endohedral guest and valence saturation, are accessible in preparative yields through a chloride‐induced disproportionation reaction of hexachlorodisilane in the presence of tri(n‐butyl)amine. X‐ray crystallography revealed that each silicon dodecahedron contains an endohedral chloride ion that imparts a net negative charge. Eight chloro substituents and twelve trichlorosilyl groups are attached to the surface of each cluster in a strictly regioregular arrangement, a thermodynamically preferred substitution pattern according to quantum‐chemical assessment. Our results demonstrate that the wet‐chemical self‐assembly of a complex, monodisperse Si nanostructure is possible under mild conditions starting from simple Si2 building blocks.  相似文献   

18.
The crystal structures of four substituted‐ammonium dichloride dodecachlorohexasilanes are presented. Each is crystallized with a different cation and one of the structures contains a benzene solvent molecule: bis(tetraethylammonium) dichloride dodecachlorohexasilane, 2C8H20N+·2Cl·Cl12Si6, (I), tetrabutylammonium tributylmethylammonium dichloride dodecachlorohexasilane, C16H36N+·C13H30N+·2Cl·Cl12Si6, (II), bis(tetrabutylammonium) dichloride dodecachlorohexasilane benzene disolvate, 2C16H36N+·2Cl·Cl12Si6·2C6H6, (III), and bis(benzyltriphenylphosphonium) dichloride dodecachlorohexasilane, 2C25H22P+·2Cl·Cl12Si6, (IV). In all four structures, the dodecachlorohexasilane ring is located on a crystallographic centre of inversion. The geometry of the dichloride dodecachlorohexasilanes in the different structures is almost the same, irrespective of the cocrystallized cation and solvent. However, the crystal structure of the parent dodecachlorohexasilane molecule shows that this molecule adopts a chair conformation. In (IV), the P atom and the benzyl group of the cation are disordered over two sites, with a site‐occupation factor of 0.560 (5) for the major‐occupied site.  相似文献   

19.
We previously reported the dinuclear material [FeII2(ddpp)2(NCS)4] ? 4 CH2Cl2 ( 1? 4 CH2Cl2; ddpp=2,5‐di(2′,2′′‐dipyridylamino)pyridine) and its partially desolvated analogue ( 1? CH2Cl2), which undergo two‐ and one‐step spin‐crossover (SCO) transitions, respectively. Here, we manipulate the type and degree of solvation in this system and find that either a one‐ or two‐step spin transition can be specifically targeted. The chloroform clathrate 1? 4 CHCl3 undergoes a relatively abrupt one‐step SCO, in which the two equivalent FeII sites within the dinuclear molecule crossover simultaneously. Partial desolvation of 1? 4 CHCl3 to form 1? 3 CHCl3 and 1? CHCl3 occurs through single‐crystal‐to‐single‐crystal processes (monoclinic C2/c to P21/n to P21/n) in which the two equivalent FeII sites become inequivalent sites within the dinuclear molecule of each phase. Both 1? 3 CHCl3 and 1? CHCl3 undergo one‐step spin transitions, with the former having a significantly higher SCO temperature than 1? 4 CHCl3 and the latter, and each has a broader SCO transition than 1? 4 CHCl3, attributable to the overlap of two SCO steps in each case. Further magnetic manipulation can be carried out on these materials through reversibly resolvating the partially desolvated material with chloroform to produce the original one‐step SCO, or with dichloromethane to produce a two‐step SCO reminiscent of that seen for 1? 4 CH2Cl2. Furthermore, we investigate the light‐induced excited spin state trapping (LIESST) effect on 1? 4 CH2Cl2 and 1? CH2Cl2 and observe partial LIESST activity for the former and no activity for the latter.  相似文献   

20.
Summary The syntheses ofArSiCl2H,ArSiH2Br, [ArSiCl2]2Hg,Ar 2SiH2,ArSi2Cl5, [ArSiCl2]2,ArSi2H5 and [ArSiH2]2 (Ar = 1-naphtyl) are reported. The compounds are characterized with29Si-NMR-spectorscopy and elemental analysis.
  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号