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.
Quantum Chemical Model Calculations on the Migration of Si? F Groups in Hexafluorosilicates The transport of different Si? F species was simulated using two [SiF6]2? octahedra as example. Activation barriers and charge distributions were calculated with the EHT method. Bearing in mind the structure of cubic hexafluorosilicates calculations were carried out on the migration both along an edge and along a (110) face of the elementary cell. At first [SiF2]2+ and SiF4 groups were removed from an [Si2F12]4? unit to produce a surface vacancy. During a second step planar SiF4 groups were moved to the neighbouring lattice position. A diffusion of planar SiF4 is favoured, if the electrostatic interaction between moved and fixed fluorine atoms is as small as possible.  相似文献   

2.
SCF-XαSW Calculations of Small Si-F Compounds SCF-Xα-SW calculations have been used to compare the bonding situations between the tetrahedral SiF4 molecule (Td), the hexafluorosilicate anion SiF62? (Oh), and the hypothetical planar SiF4 molecule (D4h). The size of the atomic spheres has been determined using both the covalent atomic radii of Slater [11] and the charge and atomic number radii according to Norman [12]. The molecular orbital analysis has been performed for the valence levels of the three molecules and the charge distributions have been calculated. Ab initio-3-21G calculations have been carried out to optimize the bond lengths of the Si-F species and to compare the results with former performed CNDO/2 calculations.  相似文献   

3.
Fluosilicic acid reacts with solutions of N,N-di-tert-butylurea (DTBU) in methanol or acetone to form crystalline compounds 2DTBU ? H2SiF6 and 2DTBU ? H2SiF6 ? Me2CO, which were characterized by the IR and 19F NMR spectra and mass spectroscopy supplemented by theoretical calculations. According to the data of IR and 19F NMR spectra, the complexes are hexafluorosilicates of O-protonated DTBU. They undergo hydrolysis in organic media with water traces; their solubility in water is very low (0.10 and 0.14 wt %, respectively). In the DTBU structure, two independent ligand molecules are joined by hydrogen bonds NH?O(N?O) 2.888(5)–2.944(5) Å).  相似文献   

4.
On the Possibility of Planar Tetracoordination of the Fourth Main Group Elements (C, Si, Ge) Semiempirical MO calculations (CNDO/2, EHT) have been used to examine the ability of silicon, carbon, and germanium to form planar tetracoordinated compounds. The calculations have been performed for the tetrahedral ground state structures (Td – symmetry) as well as for the artificial planar structures (D4h – symmetry) of the compounds CH4, SiH4, GeH4, and CF4, SiF4, GeF4. A comparison between the tetrahedral and planar species showed, that all planar species are less stable. Furthermore they have larger bond distances and their bonds are stronger polarized. The possibility of the examined compounds to form planar structures increases with growing atomic number of the central atom and with increasing electronegativity of the substituents.  相似文献   

5.
6.
Isomerisation of the Cyclotrisilazane System — Lithium Salts and Contraction Products 2,2,4,4,6,6-Hexamethyl- and 2,2,4,4,6,6-hexamethyl-1-trimethylsilyl-cyclotrisilazane ( 1, 2 ) react with n-C4H9Li to give the lithium salts 3 and 4 . At 30°C 4 isomerizes in solution to the cyclodisilazane 5 within 15 h. 4 reacts with Me2SiF2 to the substituted compound 6 , whose Li salt contracts yielding the coupled product 7 . 1,3-Bis(fluorodimethylsilyl)-2,2,4,4,6,6-hexamethylcyclotrisilazane isomerizes to the Li salt of the corresponding cyclodisilazane, which reacts with the half-molar quantity of SiF4 to the Si? N? Si? N? Si bridged cyclodisilazane dimer 8 . The tendency of contraction of 4 is discussed on the basis of its crystal structure.  相似文献   

7.
Formation of Organosilicon Compounds. 70. Reactions of Si-fluorinated 1,3,5-Trisilapentanes with CH3MgCl and LiCH3 F3Si? CCl2? SiF2? CH2? SiF3 3 reacts with meMgCl. (me = Ch3 starting with a Si-methylation and not with a C-metallation as in the corresponding Si- and C-chlorinated compounds, e. g. (Cl3Si? CCl2)2SiCl2 [2]. A CCl-hydrogenation is observed too, which in the case of F3Si? CCl2? SiF2? CHCl? SiF3 4 gives meS3Si? CCl2? Sime2? CH2? Sime3. (F3Si? CCl2)2 5 reacts with meMgCl to form preferentially 1,2-Disilapropanes by cleaving a Si? Cbond. The isolation of F3Si? CCl2H and meF2Si? CCl2? SiF2me allows to locate the bond where 5 is cleaved at the beginning of the reaction. With meLi 5 reacts to form mainly me3Si? C?C? Sime3, showing that in the reaction of meLi, being a stronger reagent than meMgCl, and 5 a C-metallation occurs, following the same mechanism as in the reaction with (Cl3Si? CCl2)2)SiCl2 [2]. The reaction conditions for the synthesis of Si-fluroinated and C-chlorinated 1,3,5-Trisilapentanes in a 0.1 mol scale are reported. N.m.r. data of all investigated compounds are tabulated.  相似文献   

8.
The relative product ion intensities from the electron-transfer reactions between SiF2+3 and the rare gases neon, argon, krypton and xenon have been rationalized using a combination of ab initio electronic structure calculations and Landau-Zener reaction window theory. The calculations show that the experimentally observed products derived from the dication (SiF+3, SiF+2 and SiF+) require the ions in the dication beam to be present in three different electronic states. The predicted and experimental product ion distributions, given this dication energy distribution, are in very close agreement. The combined computational approach adopted in this study is valuable for large molecular systems where the reactant molecule has several degrees of freedom and adopts markedly different equilibrium geometries depending on the degree of ionisation.  相似文献   

9.
The oxidative addition of BF3 to a platinum(0) bis(phosphine) complex [Pt(PMe3)2] ( 1 ) was investigated by density functional calculations. Both the cis and trans pathways for the oxidative addition of BF3 to 1 are endergonic (ΔG°=26.8 and 35.7 kcal mol?1, respectively) and require large Gibbs activation energies (ΔG°=56.3 and 38.9 kcal mol?1, respectively). A second borane plays crucial roles in accelerating the activation; the trans oxidative addition of BF3 to 1 in the presence of a second BF3 molecule occurs with ΔG° and ΔG° values of 10.1 and ?4.7 kcal mol?1, respectively. ΔG° becomes very small and ΔG° becomes negative. A charge transfer (CT), F→BF3, occurs from the dissociating fluoride to the second non‐coordinated BF3. This CT interaction stabilizes both the transition state and the product. The B?F σ‐bond cleavage of BF2ArF (ArF=3,5‐bis(trifluoromethyl)phenyl) and the B?Cl σ‐bond cleavage of BCl3 by 1 are accelerated by the participation of the second borane. The calculations predict that trans oxidative addition of SiF4 to 1 easily occurs in the presence of a second SiF4 molecule via the formation of a hypervalent Si species.  相似文献   

10.
Si?F bond cleavage of fluoro‐silanes was achieved by transition‐metal complexes under mild and neutral conditions. The Iridium‐hydride complex [Ir(H)(CO)(PPh3)3] was found to readily break the Si?F bond of the diphosphine‐ difluorosilane {(o‐Ph2P)C6H4}2Si(F)2 to afford a silyl complex [{[o‐(iPh2P)C6H4]2(F)Si}Ir(CO)(PPh3)] and HF. Density functional theory calculations disclose a reaction mechanism in which a hypervalent silicon species with a dative Ir→Si interaction plays a crucial role. The Ir→Si interaction changes the character of the H on the Ir from hydridic to protic, and makes the F on Si more anionic, leading to the formation of Hδ+???Fδ? interaction. Then the Si?F and Ir?H bonds are readily broken to afford the silyl complex and HF through σ‐bond metathesis. Furthermore, the analogous rhodium complex [Rh(H)(CO)(PPh3)3] was found to promote the cleavage of the Si?F bond of the triphosphine‐monofluorosilane {(o‐Ph2P)C6H4}3Si(F) even at ambient temperature.  相似文献   

11.
Deprotonation of aminophosphaalkenes (RMe2Si)2C?PN(H)(R′) (R=Me, iPr; R′=tBu, 1‐adamantyl (1‐Ada), 2,4,6‐tBu3C6H2 (Mes*)) followed by reactions of the corresponding Li salts Li[(RMe2Si)2C?P(M)(R′)] with one equivalent of the corresponding P‐chlorophosphaalkenes (RMe2Si)2C?PCl provides bisphosphaalkenes (2,4‐diphospha‐3‐azapentadienes) [(RMe2Si)2C?P]2NR′. The thermally unstable tert‐butyliminobisphosphaalkene [(Me3Si)2C?P]2NtBu ( 4 a ) undergoes isomerisation reactions by Me3Si‐group migration that lead to mixtures of four‐membered heterocyles, but in the presence of an excess amount of (Me3Si)2C?PCl, 4 a furnishes an azatriphosphabicyclohexene C3(SiMe3)5P3NtBu ( 5 ) that gave red single crystals. Compound 5 contains a diphosphirane ring condensed with an azatriphospholene system that exhibits an endocylic P?C double bond and an exocyclic ylidic P(+)? C(?)(SiMe3)2 unit. Using the bulkier iPrMe2Si substituents at three‐coordinated carbon leads to slightly enhanced thermal stability of 2,4‐diphospha‐3‐azapentadienes [(iPrMe2Si)2C?P]2NR′ (R′=tBu: 4 b ; R′=1‐Ada: 8 ). According to a low‐temperature crystal‐structure determination, 8 adopts a non‐planar structure with two distinctly differently oriented P?C sites, but 31P NMR spectra in solution exhibit singlet signals. 31P NMR spectra also reveal that bulky Mes* groups (Mes*=2,4,6‐tBu3C6H2) at the central imino function lead to mixtures of symmetric and unsymmetric rotamers, thus implying hindered rotation around the P? N bonds in persistent compounds [(RMe2Si)2C?P]2NMes* ( 11 a , 11 b ). DFT calculations for the parent molecule [(H3Si)2C?P]2NCH3 suggest that the non‐planar distortion of compound 8 will have steric grounds.  相似文献   

12.
The mechanism of the gas‐phase reactions of SiHn+ (n = 1,2) with NF3 were investigated by ab initio calculations at the MP2 and CAS‐MCSCF level of theory. In the reaction of SiH+, the kinetically relevant intermediates are the two isomeric forms of fluorine‐coordinated intermediate HSi‐F‐NF2+. These species arise from the exoergic attack of SiH+ to one of the F atoms of NF3 and undergo two competitive processes, namely an isomerization and subsequent dissociation into SiF+ + HNF2, and a singlet‐triplet crossing so to form the spin‐forbidden products HSiF+ + NF2. The reaction of SiH2+ with NF3 involves instead the concomitant formation of the nitrogen‐coordinated complex H2Si‐NF3+ and of the fluorine‐coordinated complex H2Si‐F‐NF2+. The latter isomer directly dissociates into NF2+ + H2SiF, whereas the former species preferably undergoes the passage through a conical intersection point so to form a H2SiF‐NF2+ isomer, which eventually dissociates into H2SiF+ and NF2. © 2012 Wiley Periodicals, Inc.  相似文献   

13.
F3Si? PH2 F3Si? PH2 is formed according to eq. (1) in 60% yield, whereas Br2SiF2 and Br3SiF do not yield the corresponding Silylphosphines for reasons of dismutation. The nmr- and ir-data are reported. F3Si? PH2 is not cleaved by HE(CF3)2 (E = P, As), however, it reacts with JE(CF3)2 in a manner similar to H3Si? PH2 as is shown by eq. (2) Due to the lower basicity of the P-atom in F3Si? PH2 this compound is more stable than H3Si? PH2. The adducts formed with BCl3, BBr3, and AlCl3 are less stable than those of H3Si? PH2. Therefore H3Si? PH2 is cleaved more readily than F3Si? PH2.  相似文献   

14.
At beginning thermal decomposition K2[SiF6] loses SiF4-planes from [SiF6]2?-octahedrons, which has been proved by x-ray-diffraction [1], [2]. Analogous disorder structures are supposed to be present with all solids having complex ions including carbonates, sulfates and others. The result is a high reactivity at this spots. Another reactive form in hexefluorosilicates is represented by mobile SiF-species, perhaps SiF3+. The reactivity is shown by heterogenous reactions with CHCl3 and by solid-solid reactions for instance with halides, oxides etc. As an example corundum (α-Al2O3) reacts at 600°C giving K3 AlF6 and KAlSiO4 [3].  相似文献   

15.
A good understanding of gas‐phase fragmentation chemistry of peptides is important for accurate protein identification. Additional product ions obtained by sodiated peptides can provide useful sequence information supplementary to protonated peptides and improve protein identification. In this work, we first demonstrate that the sodiated a3 ions are abundant in the tandem mass spectra of sodium‐cationized peptides although observations of a3 ions have rarely been reported in protonated peptides. Quantum chemical calculations combined with tandem mass spectrometry are used to investigate this phenomenon by using a model tetrapeptide GGAG. Our results reveal that the most stable [a3 + Na ? H]+ ion is present as a bidentate linear structure in which the sodium cation coordinates to the two backbone carbonyl oxygen atoms. Due to structural inflexibility, further fragmentation of the [a3 + Na ? H]+ ion needs to overcome several relatively high energetic barriers to form [b2 + Na ? H]+ ion with a diketopiperazine structure. As a result, low abundance of [b2 + Na ? H]+ ion is detected at relatively high collision energy. In addition, our computational data also indicate that the common oxazolone pathway to generate [b2 + Na ? H]+ from the [a3 + Na ? H]+ ion is unlikely. The present work provides a mechanistic insight into how a sodium ion affects the fragmentation behaviors of peptides. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

16.
The structure and properties of fluorosilicon polymer (fluorosil) formed by the reaction of phenyltrifluorosilane with aliphatic alcohols have been studied by the methods of IR, 19F, 29Si NMR spectroscopy, high temperature mass-spectrometry, derivatography and atom emission analysis. Due to its high reactivity, this polymer readily reacts with glass of the reaction vessel extracting the ions of all metals entering into its composition. Fluorosil formed in a quartz, teflon or polypropylene reactors is characterized by low stability and is slowly decomposed to SiF4 and SiO2. Apparently, fluorosil is the product of incorporation of SiF4 into the SiO2 matrix.  相似文献   

17.
The structure of the complex (o-CH3C6H4NH3)2SiF6 was determined by X-ray diffraction. In the ionic structure of the complex, the SiF 6 2? anions (Si-F, 1.595(9)-1.683(17)Å) and the o-CH3C6H4NH 3 + cations are combined by NH?F hydrogen bonds (N?F 2.757(10)-3.25(2) Å). The components of the structure are combined into a layer whose central part is formed by the SiF 6 2? anions and the outer hydrophobic surfaces are formed by the aromatic rings of the cations.  相似文献   

18.
19.
The stability of noble gas (Ng)‐bound SiH3+ clusters is explored by ab initio computations. Owing to a high positive charge (+1.53 e?), the Si center of SiH3+ can bind two Ng atoms. However, the Si?Ng dissociation energy for the first Ng atom is considerably larger than that for the second one. As we go down group 18, the dissociation energy gradually increases, and the largest value is observed for the case of Rn. For NgSiH3+ clusters, the Ar–Rn dissociation processes are endergonic at room temperature. For He and Ne, a much lower temperature is required for it to be viable. The formation of Ng2SiH3+ clusters is also feasible, particularly for the heavier members and at low temperature. To shed light on the nature of Si?Ng bonding, natural population analysis, Wiberg bond indices computations, electron‐density analysis, and energy‐decomposition analysis were performed. Electron transfer from the Ng centers to the electropositive Si center occurs only to a small extent for the lighter Ng atoms and to a somewhat greater extent for the heavier analogues. The Si?Xe/Rn bonds can be termed covalent bonds, whereas the Si?He/Ne bonds are noncovalent. The Si?Ar/Kr bonds possess some degree of covalent character, as they are borderline cases. Contributions from polarization and charge transfer and exchange are key terms in forming Si?Ng bonds. We also studied the effect of substituting the H atoms of SiH3+ by halide groups (?X) on the Ng binding ability. SiF3+ showed enhanced Ng binding ability, whereas SiCl3+ and SiBr3+ showed a lower ability to bind Ng than SiH3+. A compromise originates from the dual play of the inductive effect of the ?X groups and X→Si π backbonding (pz–pz interaction).  相似文献   

20.
In bis(2‐carboxypyridinium) hexafluorosilicate, 2C6H6NO2+·SiF62−, (I), and bis(2‐carboxyquinolinium) hexafluorosilicate dihydrate, 2C10H8NO2+·SiF62−·2H2O, (II), the Si atoms of the anions reside on crystallographic centres of inversion. Primary inter‐ion interactions in (I) occur via strong N—H...F and O—H...F hydrogen bonds, generating corrugated layers incorporating [SiF6]2− anions as four‐connected net nodes and organic cations as simple links in between. In (II), a set of strong N—H...F, O—H...O and O—H...F hydrogen bonds, involving water molecules, gives a three‐dimensional heterocoordinated rutile‐like framework that integrates [SiF6]2− anions as six‐connected and water molecules as three‐connected nodes. The carboxyl groups of the cation are hydrogen bonded to the water molecule [O...O = 2.5533 (13) Å], while the N—H group supports direct bonding to the anion [N...F = 2.7061 (12) Å].  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