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 730 毫秒
1.
2.
3.
A multilevel approach that combines high‐level ab initio quantum chemical methods applied to a molecular model of a single, strain‐free Si O Si bridge has been used to derive accurate energetics for Si O bond cleavage. The calculated Si O bond dissociation energy and the activation energy for water‐assisted Si O bond cleavage of 624 and 163 kJ mol−1, respectively, are in excellent agreement with values derived recently from experimental data. In addition, the activation energy for H2O‐assisted Si O bond cleavage is found virtually independent of the amount of water molecules in the vicinity of the reaction site. The estimated reaction energy for this process including zero‐point vibrational contribution is in the range of −5 to 19 kJ mol−1. © 2017 Wiley Periodicals, Inc.  相似文献   

4.
Poly[(methylsilylene)ethynylene] ( 1 ) and poly[2,5-thiophenediyl(methylsilylene)] ( 2 ) are prepared in good yields. Thermogravimetric analysis and differential scanning calorimetry indicate the Si H group to increase the pyrolysis residue yields. Gelation was achieved from polymer 1 to get improved preceramic materials. © 1998 John Wiley & Sons, Inc. J. Polym. Sci. A Polym. Chem. 36: 2275–2282, 1998  相似文献   

5.
The reaction of two‐coordinated (trimethylsilylamino)phosphines (Me3Si)2N PE SiMe3 1 (E = N) and 2 (E = CH) with hydroxycarbonyl compounds proceeded with four‐ or five‐member heterocyclization to yield derivatives of oxaphosphetane, oxaphospholanes, and oxaphospholes. The reaction rate depends on the structure of hydroxyketones as well as on the type of the two‐coordinated phosphorus compound in accordance with the polarity of the P=N and P=C bonds. Thus, reaction was completed in 30 min in the case of the ortho carbonyl phenoxy derivatives with the phosphine 1 , but required 2 h in the case of the alkyl hydroxy carbonyls. All reactions with the phosphine 2 took about 24 h. © 2004 Wiley Periodicals, Inc. Heteroatom Chem 15:413–417, 2004; Published online in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/hc.20033  相似文献   

6.
Nitride Sulfide Chlorides of the Lanthanides. II. The Composition M6N3S4Cl (M = La? Nd) The oxidation of the “light” lanthanides (M = La? Nd) with sulfur and NaN3 in the presence of the chlorides MCl3 yields chlorine-poor nitride sulfide chlorides with the composition M6N3S4Cl when appropriate molar ratios of the reactants are used. Additional NaCl as a flux secures complete and fast reactions (7 d) at 850°C in evacuated silica vessels as well as single-crystalline products (red-brown needles). The crystal structure was determined from X-ray single crystal data for the limiting representatives La6N3S4Cl (orthorhombic, Pnma (no. 62), Z = 4, a = 1159.7(4), b = 410.95(7), c = 2756.8(9)pm, R = 0.030, Rw = 0.027) and Nd6N3S4Cl (a = 1137.1(3), b = 399.34(6), c = 2687.6(9)pm, R = 0.034, Rw = 0.033). Guinier powder data revealed the cerium and praseodymium analogues to be isotypic. The crystal structure exhibits two different chains of connected [NM4] tetrahedra which are commensurate in translation. Six crystallographically different M3+ are present, two of them (M1 and M2) build up the chain [(N1)(M1) · (M2)]3+ together with (N1)3? by cis-edge connection of tetrahedra. The four remainders (M3? M6) arrange as pairs [N2M6] of edge-shared [NM4] tetrahedra with (N2)3? and (N3)3? which are further connected via four vertices to form the [(M5)(N-2){(M3)(1+1)/(1+1)(M4)(1+1)/(1+1))}e(N3)(M6)]6+ double chain. Bundled along [010] like a closest packing of rods, both types of chains are held together by five crystallographically different but by X-ray diffraction indistinguishable anions S2? (S1? S4) and Cl? adjusting the charge balance in a molar ratio of 4:1.  相似文献   

7.
Oxidative addition plays a major role in transition‐metal catalysis, but this elementary step remains very elusive in gold chemistry. It is now revealed that in the presence of GaCl3, phosphine gold chlorides promote the oxidative addition of disilanes at low temperature. The ensuing bis(silyl) gold(III) complexes were characterized by quantitative 31P and 29Si NMR spectroscopy. Their structures (distorted Y shape) and the reaction profile of σ(Si Si) bond activation were analyzed by DFT calculations. These results provide evidence for the intermolecular oxidative addition of σ(Si Si) bonds to gold and open promising perspectives for the development of new gold‐catalyzed redox transformations.  相似文献   

8.
The structures and energetics of eight substituted bis(thiocarbonyl)disulfides (RCS2)2, their associated radicals RCS2., and their coordination compounds with a lithium cation have been studied at the G3X(MP2) level of theory for R=H, Me, F, Cl, OMe, SMe, NMe2, and PMe2. The effects of substituents on the dissociation of (RCS2)2 to RCS2. were analyzed using isodesmic stabilization reactions. Electron‐donating groups with an unshared pair of electrons have a pronounced stabilization effect on both (RCS2)2 and RCS2.. The S? S bond dissociation enthalpy of tetramethylthiuram disulfide (TMTD, R=NMe2) is the lowest in the above series (155 kJ mol?1), attributed to the particular stability of the formed Me2NCS2. radical. Both (RCS2)2 and the fragmented radicals RCS2. form stable chelate complexes with a Li+ cation. The S? S homolytic bond cleavage in (RCS2)2 is facilitated by the reaction [Li(RCS2)2]++Li+→2 [Li(RCS2)].+. Three other substituted bis(thiocarbonyl) disulfides with the unconventional substituents R=OSF5, Gu1, and Gu2 have been explored to find suitable alternative rubber vulcanization accelerators. Bis(thiocarbonyl)disulfide with a guanidine‐type substituent, (Gu1CS2)2, is predicted to be an effective accelerator in sulfur vulcanization of rubber. Compared to TMTD, (Gu1CS2)2 is calculated to have a lower bond dissociation enthalpy and smaller associated barrier for the S? S homolysis.  相似文献   

9.
The reactivity of the cubane‐type rare‐earth methylidene complex [Cp′Lu(μ3‐CH2)]4 ( 1 , Cp′=C5Me4SiMe3) with various unsaturated electrophiles was investigated. The reaction of 1 with CO (1 atm) at room temperature gave the bis(ketene dianion)/dimethylidene complex [Cp′4Lu43‐CH2)232‐O‐C?CH2)2] ( 2 ) in 86 % yield through the insertion of two molecules of CO into two of the four lutetium–methylidene units. In the reaction with the sterically demanding N,N‐diisopropylcarbodiimide at 60 °C, only one of the four methylidene units in 1 reacted with one molecule of the carbodiimide substrate to give the mono(ethylene diamido)/trimethylidene complex [Cp′4Lu43‐CH2)3{iPrNC(=CH2)NiPr}] ( 3 ) in 83 % yield. Similarly, the reaction of 1 with phenyl isothiocyanate gave the ethylene amido thiolate/trimethylidene complex [Cp′4Lu43‐CH2)3{PhNC(S)=CH2}] ( 4 ). In the case of phenyl isocyanate, two of the four methylidene units in 1 reacted with four molecules of the substrate at ambient temperature to give the malonodiimidate/dimethylidene complex [Cp′4Lu43‐CH2)2{PhN=C(O)CH2(O)C?NPh}2] ( 5 ) in 87 % yield. In this reaction, each of the two lutetium–methylidene bonds per methylidene unit inserted one molecule of phenyl isocyanate. All the products have been fully characterized by NMR spectroscopy, X‐ray diffraction, and microelemental analyses.  相似文献   

10.
Bridging or pendant? Palladium and rhodium complexes deriving from an ambiphilic phosphine–borane ligand are shown to adopt a bridging P→M? Cl→B coordination mode in the solid state. DFT calculations provide more insight into the Cl→B interaction and suggest the possible interconversion of the bridging and B‐pendant forms in solution.

  相似文献   


11.
To improve the emission and excited‐state properties of luminescent cyanometalates, new classes of highly solvatochromic luminescent cyanoruthenium(II) and cyanoruthenate(II) complexes of the general formulae [Ru(PR3)2(CN)2($\widehat{NN}$ )] and K[Ru(PR3)(CN)3($\widehat{NN}$ )], respectively, were developed. These complexes could be readily synthesized through the ligand‐substitution reaction of K2[Ru(CN)4(PR3)2] with a diimine ligand. The geometrical isomerism of these complexes was characterized by using various spectroscopic techniques. Their photophysical properties, solvatochromism, and electrochemistry have also been investigated. Our detailed study showed that many of these complexes exhibited extremely environmentally sensitive emissions and significantly improved emission quantum efficiencies and lifetimes compared with the well‐studied tetracyanoruthenate systems.  相似文献   

12.
The IR spectrum of Si3H8+ ions produced in a supersonic plasma molecular beam expansion of SiH4, He, and Ar is inferred from photodissociation of cold Si3H8+–Ar complexes. Vibrational analysis of the spectrum is consistent with a Si3H8+ structure ( 2+ ) obtained by a barrierless addition reaction of SiH4 to the disilene ion (H2Si?SiH2+) in the silane plasma. In this structure, one of the electronegative H atoms of SiH4 donates electron density into the partially filled electrophilic π orbital of the disilene cation. The resulting asymmetric Si? H? Si bridge of the 2+ isomer with a bond energy of approximately 60 kJ mol?1 is characteristic for a weak three‐center two‐electron bond, which is identified by its strongly IR active asymmetric Si? H? Si stretching fundamental at about 1765 cm?1. The observed 2+ isomer is calculated to be only a few kJ mol?1 less stable than the global minimum structure of Si3H8+ ( 1+ ), which is derived from vertical ionization of trisilane. Although more stable, 1+ is not detected in the measured IR spectrum of Si3H8+–Ar, and its lower abundance in the supersonic plasma is rationalized by the production mechanism of Si3H8+ in the silane plasma, in which a high barrier between 2+ and 1+ prevents the efficient formation of 1+ . The potential energy surface of Si3H8+ is characterized in some detail by quantum chemical calculations. The structural, vibrational, electronic and energetic properties as well as the chemical bonding mechanism are investigated for a variety of low‐energy Si3H8+ isomers and their fragments. The weak intermolecular bonds of the Ar ligands in the Si3H8+–Ar isomers arise from dispersion and induction forces and induce only a minor perturbation of the bare Si3H8+ ions. Comparison with the potential energy surface of C3H8+ reveals the differences between the silicon and carbon species.  相似文献   

13.
Electroactive fused ethylenedithio? tetrathiafulvalene? [4]helicene and ‐[6]helicenes have been synthesized through a strategy that involved the preparation of 2,3‐dibromo‐helicene derivatives as intermediates. The dihedral angles between the terminal helicenes, as determined by single‐crystal X‐ray analysis, are 22.7° and 50.7° for the [4]helicene and [6]helicene, respectively. Their solid‐state architectures show interplay between S???S and π???π intermolecular interactions. The chiroptical properties of the enantiopure EDT? TTF? [6]helicene derivatives have been investigated and supported by TDDFT calculations. Remarkable redox switching of the circular dichroism (CD) signal between the neutral and radical‐cation species has been achieved.  相似文献   

14.
A “niche” topic in the past decade, the asymmetric C? H bond activation has been attracting growing interest over the last few years. Particularly significant advances have been achieved in the field of direct, stereoselective transformations of C(sp2)? H bonds. This Concept article intends to showcase different types of asymmetric C(sp2)? H bond activation reactions, emphasising both the nature of the stereo‐discriminating step and the variability of valuable scaffolds that could be rapidly constructed by means of such strategies.  相似文献   

15.
A computational study of the intramolecular pnicogen bond in PHF? (CH2)n? PHF (n=2–6) systems was carried out. For each compound, two different conformations, (R,R) and (R,S), were considered on the basis of the chirality of the phosphine groups. The characteristics of the closed conformers, in which the pnicogen interaction occurs, were compared with those of the extended conformer. In several cases, the closed conformations are more stable than the extended conformations. The calculated interaction energies of the pnicogen contact, by means of isodesmic reactions, provide values between ?3.4 and ?26.0 kJ mol?1. Atoms in molecules and electron localization function analysis of the electron density showed that the systems in the closed conformations with short P ??? P distances have a partial covalent character in this interaction. The calculated absolute chemical shieldings of the P atoms showed an exponential relationship with the P ??? P distance. In addition, a search in the Cambridge crystallographic database was carried out to detect those compounds with a potential intramolecular pnicogen bond in the solid phase.  相似文献   

16.
N-Silylation and Si? O Bond Splitting at the Reaction of Lithiated Siloxy-silylamino-silanes with Chlorotrimethylsilane Lithiated Siloxy-silylamino-silanes were allowed to react in tetrahydrofurane (THF) and in n-octane (favoured) and n-hexane, resp., with chlorotrimethylsilane. The monoamide (Me3SiO)Me2Si(NLiSiMe3) gives in THF and in n-octane the N-substitution product (Me3SiO)Me2Si · [N(SiMe3)2] 1 , the diamide (Me3SiO)MeSi(NLiSiMe3)2 only in THF the N-substitution products (Me3SiO)MeSi[N(SiMe3)2]2 2 (main product) and (Me3SiO)MeSi[N(SiMe3)2](NHSiMe3) 3 . In n-octane the diamide reacts mainly under Si? O bond splitting. The cyclodisilazane [(Me3SiNH)MeSi? NSiMe3]2 6 is obtained as the main product. Byproducts are 2, 3 and the tris(trimethylsilylamino) substituted disilazane (Me3SiO)(Me3SiNH)MeSi? N · (SiMe3)? SiMe(NHSiMe3)2 7 . The triamide (Me3SiO)Si · (NLiSiMe3)3 reacts under Si? O and Si? N bond splitting in n-octane as well as in THF. The cyclodisilazanes [(Me3SiNH)2 · Si? NSiMe3]2 10 and ( 11 : R = Me3SiNH, 12 : R = (Me3Si)2N) are formed. in THF furthermore the N-substitution products (Me3SiO)Si[N(SiMe3)2] · (NHSiMe3)2 4 and (Me3SiO)Si[N(SiMe3)2]2(NHSiMe3) 5 . The Si? O bond splitting occurs in boiling n-octane also in absence of the chlorotrimethylsilane. An amide solution of (Me3SiO)MeSi(NHSiMe3)2 with n-butyllithium in the molar ratio 1 : 1 leads in n-octane and n-hexane to 6 and 7 , in THF to 3 . The amide solutions of (Me3SiO)Si · (NHSiMe3)3 with n-butyllithium the molar ratio 1 : 1 and 1 : 2 give in THF 4 and 5 , respectively.  相似文献   

17.
A dimer of thioxo-N-t-butylimino(trimethylsiloxy)-phosphorane 5 has been prepared by reaction of tris(trimethylsilyl) phosphine with N-sulfinyl-N-tert-butylamine. The structure of 5 has been confimed by X-ray analysis data. 1-Aza-2-thia-3-phosphaallene 1 , thiaphosphaziridine 3 , iminophosphine P-sulfide 4 are postulated as intermediates of the reaction studied.  相似文献   

18.
The addition of NO (0 to 400ppm) to mixtures of H2 (ca. 1%) and O2 (0.7 to 22%) has been studied over the temperature range 700 to 825 K, in a flow reactor at atmospheric pressure. The overall effect of NO is to promote the oxidation of H2 but high concentrations of O2 actually inhibit the NO-promoted oxidation of H2. A detailed kinetic mechanism has been constructed and found to describe the experimental observations. The promotion of the oxidation of H2 arises through the catalytic cycle The ability of R.34 to reactivate chains normally terminated by the formation of HO2 is a key feature of this system. The predictions are highly sensitive to the rate of the reaction R.5 and the rate constants for this reaction is the only adjustable parameter required in the model. The value of k5,N2 found to describe all the results has an absolute uncertainty <35%. The uncertainty relative to other important rate constants in the H2? O2 system is less than 10%. © 1995 John Wiley & Sons, Inc.  相似文献   

19.
The reactivity of disulfide and diselenide derivatives towards F? and CN? nucleophiles has been investigated by means of B3PW91/6‐311+G(2df,p) calculations. This theoretical survey shows that these processes, in contrast with the generally accepted view of disulfide and diselenide linkages, do not always lead to S? S or Se? Se bond cleavage. In fact, S? S or Se? Se bond fission is the most favorable process only when the substituents attached to the S or the Se atoms are not very electronegative. Highly electronegative substituents (X) strongly favor S? X bond fission. This significant difference in the observed reactivity patterns is directly related to the change in the nature of the LUMO orbital of the disulfide or diselenide derivative as the electronegativity of the substituents increases. For weakly electronegative substituents, the LUMO is a σ‐type S? S (or Se? Se) antibonding orbital, but as the electronegativity of the substituents increases the π‐type S? X antibonding orbital stabilizes and becomes the LUMO. The observed reactivity also changes with the nature of the nucleophile and with the S or Se atom that undergoes the nucleophilic attack in asymmetric disulfides and diselenides. The activation strain model provides interesting insights into these processes. There are significant similarities between the reactivity of disulfides and diselenides, although some dissimilarities are also observed, usually related to the different interaction energies between the fragments produced in the fragmentation process.  相似文献   

20.
Nitride Sulfide Chlorides of the Lanthanides. I. The Composition M4NS3Cl3 (M = La Nd) The oxidation of the „light”︁ lanthanides (M = La Nd) with sulfur and NaN3 the presence of the chlorides MCl3 results in the formation of the first lanthanide nitride sulfide chlorides M4NS3Cl3 when appropriate molar ratios of the reactants are used. The addition of some NaCl (or an excess of MCl3) as a flux secures complete and fast reaction (7 d) at 850°C in evacuated silica vessels as well as single-crystalline products. Since these nitride sulfide chlorides (fine transparent needles) are not sensitive against hydrolysis, the surplus chloride can be removed easily with water. The crystal structure was determined from X-ray single crystal data for the example of La4NS3Cl3 (hexagonal, P63mc (no. 186), Z = 2, a = 941.40(3), c = 700.36(3) pm, R = 0.026, Rw = 0.021) and the nitride sulfide chlorides M4NS3Cl3 with M = Ce Nd proved to be isostructural from Guinier powder data. According to their Ba3OCl6-analogue structure, two crystallographically different M3+ cations are present (CN(M1) = 10, CN(M2) = 8). „Isolated”︁ tetrahedra [(N3−)(M3+)4] build up the Mayn structural feature according to [NM4]S3Cl3. They are hexagonally closest packed and interconnected via the crystallographically different but by X-ray diffraction indistinguishable anions S2− and Cl, which take care for charge neutrality.  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