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 15 毫秒
1.
The synthesis and crystal structures of two dinuclear titanocene hydride complexes are reported. Both complexes, namely bis(η5‐(di‐para‐tolylmethyl)cyclopentadienyl)titanium hydride dimer, [(η5‐C20H19)2Ti(μ‐H)]2 ( 2a ), and bis(η5‐2‐adamantylcyclopentadienyl)‐titanium hydride dimer, [(η5‐C15H19)2Ti(μ‐H)]2 ( 2b ), are formed via activation of molecular hydrogen by the corresponding bis(η51‐pentafulvene)titanium complexes 1a and 1b at ambient temperatures and pressures in high yields. The hydride complexes 2a and 2b exhibit planar [Ti2H2] cores and, as a result of the heterolytic cleavage of molecular hydrogen, substituted Cp Ligands were formed during the reaction.  相似文献   

2.
The super acidity of the unsolvated Al(C6F5)3 enabled isolation of the elusive silane–alane complex [Si? H???Al], which was structurally characterized by spectroscopic and X‐ray diffraction methods. The Janus‐like nature of this adduct, coupled with strong silane activation, effects multifaceted frustrated‐Lewis‐pair‐type catalysis. When compared with the silane–borane system, the silane–alane system offers unique features or clear advantages in the four types of catalytic transformations examined in this study, including: ligand redistribution of tertiary silanes into secondary and quaternary silanes, polymerization of conjugated polar alkenes, hydrosilylation of unactivated alkenes, and hydrodefluorination of fluoroalkanes.  相似文献   

3.
A novel transformation involving phosphine? diazo ester zwitterions (generated from dialkyl azodicarboxylates with Ph3P) and α‐(alkoxycarbonyl)imidoyl chlorides (prepared from α‐addition of acyl chlorides to alkyl isocyanides) to afford dialkyl 2‐[3‐alkoxy‐1‐(alkylimino)‐1‐chloro‐3‐oxopropan‐2‐ylidene]hydrazine‐1,1‐dicarboxylates in moderate yields, is described.  相似文献   

4.
Lewis acid‐base adducts of the general type R2Zn(4‐tBuPy)x (R = Me 1 , iPr 2 , tBu 3 , Cp* 4 ; x = 1, 2) were obtained in high yields from reactions of ZnR2 with the Lewis base 4‐tBu‐Pyridine. Compounds 1 – 4 were characterized by multinuclear NMR (1H, 13C) and IR spectroscopy and elemental analyses, 1 and 4 also by X‐ray diffraction at single crystals.  相似文献   

5.
Several salts of protonated amines and aza‐aromatics with [AuCl4] and [AuBr4] anions contain two‐dimensional (“square”) anionic networks that display short halogen ··· halogen contacts. The Au4 quadrilaterals formed by neighboring anions of the networks are to a good approximation squares, with sides of around 7.5 Å for tetrachloridoaurates and 8 Å for tetrabromidoaurates.  相似文献   

6.
The single‐crystal X‐ray structure analysis of hexakis(2,4,6‐triisopropylphenyl)cyclotristannoxane, cyclo‐[(2,4,6‐i‐Pr3‐C6H2)2SnO]3 ( 1 ), is reported and reveals this compound to contain an almost planar six‐membered ring. Redistribution reactions of 1 with cyclo‐(t‐Bu2SnO)3 and t‐Bu2SiCl2, respectively, failed and indicate an unusual kinetic inertness of the Sn–O bonds in 1 as compared to related molecular diorganotin oxides containing less bulkier substituents. The redistribution reaction of cyclo‐(t‐Bu2SnO)3 with cyclo‐(t‐Bu2SnS)2 leads to an equilibrium involving the trimeric diorganotin oxysulphides cyclot‐Bu2Sn(OSnt‐Bu2)2S ( 2 a ) and cyclot‐Bu2Sn(SSnt‐Bu2)2O ( 2 b ).  相似文献   

7.
8.
9.
Solid solution phases Li7‐2xMgx[VN4] (0 < x ≤ 1) with varying Mg‐content are obtained as yellow microcrystalline powders from heat treatment of mixtures of VN, Li3N and Mg3N2 or from mixtures of Li7[VN4] and Mg3N2 at 1370 K in N2 atmosphere at ambient pressure. At substitution parameter values of x > 0.5 a subsequent distortion from the ideal cubic unit cell to an orthorhombic unit cell is observed. The crystal structure of Li7‐2xMgx[VN4] with x ≈ 1 was refined from neutron and X‐ray powder diffraction data (space group Pbca, No. 61, a = 963.03(3) pm, b = 958.44(3) pm, c = 951.93(2) pm, neutron pattern 14° — 156° 2θ, step non‐linear ≈ 0.0782° 2θ, No. of measured points 1816, Rp = 0.089, Rwp = 0.115, RBragg = 0.155, RF = 0.114; X‐ray pattern 10° — 98° 2θ, step 0.005° 2θ, No. of measured points 17600, Rp = 0.028, Rwp = 0.045, RBragg = 0.113, RF = 0.133, structure variables: 45). The crystal structure resembles a Li2O type superstructure with the atomic arrangement of β‐Li7[VN4] and with two crystallographic Li‐sites each substituted by Mg with statistical occupation factors of 0.5. Chemical analyses prove the composition and XAS spectroscopy at the V K‐edge support the +5 oxidation state assignment for vanadium. XAS data also support the tetrahedral coordination of vanadium by N as indicated by the structure refinements.  相似文献   

10.
Compounds of the three large cations tetramethylammonium, tetramethylphosphonium, and tetramethylarsonium with the superoxide radical anion were synthesized by either metathesis or ion exchange in liquid ammonia. They were obtained from concentrated solutions as ammoniates in the form of long needle‐shaped single crystals. [N(CH3)4]‐(O2)?3NH3 crystallizes in the monoclinic crystal system, whereas the two compounds [E(CH3)4](O2)?2NH3 (E=P, As) are isostructural and belong to the orthorhombic crystal system. The cation–anion packing in all three crystal structures is related to the sodium chloride structure. All structures contain hydrogen bonds between the ammonia molecules and between ammonia and the superoxide. The solvent of crystallization was easily released from the crystals upon complete removal of the solvent from the reaction vessel, leading to polycrystalline samples. The Raman spectra of all three solvent‐free compounds show the symmetric stretching mode of the superoxide ion at about 1123 cm?1. The desolvated [N(CH3)4](O2) was investigated by powder X‐ray diffraction, and the crystal structure was solved by ab initio simulated annealing methods by using rigid‐body models of the constituent molecular ions. The superoxide ion shows rotational disorder. The magnetic susceptibility of tetramethylammonium superoxide follows the Curie–Weiss law with a high‐temperature effective magnetic moment of 1.66(3) μB and a paramagnetic Curie temperature of Θ=?13(6) K. Complementary electron paramagnetic resonance spectroscopy revealed that the average g factor is temperature‐dependent. It decreased from 2.15 at 10 K to 1.66 at 100 K, possibly due to the onset of rotational motion of the superoxide ion and in accordance with the lower‐than‐expected effective magnetic moment.  相似文献   

11.
Cations derived by protonation of the ligand title compound (L1) have been structurally characterized in their di‐ and tetra‐ protonated forms in the salts [H2L1][ClO4]2·2H2O and [H4L1][ZnCl4]2·4H2O. In both structures, one half of the formula unit comprises the asymmetric unit of the structure, the macrocycle being centrosymmetric, with the two macrocycles adopting similar conformations. In both salts, a pair of diagonally opposed macrocyclic secondary amine groups are protonated; in the [H4L1]4+ salt, the additional pair of protons are accommodated on the exocyclic pendant amine groups. The dispositions of the pendent amines differ between the two structures, being ‘equatorial’ with respect to the macrocyclic ring in the [H2L1]2+ salt, and ‘axial’ in the [H4L1]4+ salt. In other structurally characterized compounds containing [H4L1]4+ the equatorial disposition was found in the ferricyanide adduct, while in the tetraperchlorate salt the axial disposition was identified. The differences in disposition of the exocyclic groups are ascribed to the extensive H‐bonding in the lattices.  相似文献   

12.
The reaction of germanium(II)‐bis(2‐methoxyphenyl)methoxide with methanesulfonic acid provides the germanium(II) sulfonate Ge(CH3SO3)2 ( 1 ), which was characterized by X‐ray diffraction, elemental analysis, NMR spectroscopy, and IR spectroscopy. The decomposition process of 1 was investigated by thermal gravimetric analysis (TGA) and temperature‐dependent X‐ray powder diffraction (PXRD) and both are consistent with the formation of GeO2 as major final product. Single crystal X‐ray diffraction at 110 K revealed the chiral tetragonal space group P41212 and formation of a three‐dimensional (3D) coordination network solid. The 3D network is composed of interconnected twenty four‐membered rings comprising bridging methanesulfonate groups, which link the germanium atoms.  相似文献   

13.
A high‐yield synthesis toward 5,5′‐bis(silyl)‐functionalized 3,3′‐dibromo‐2,2′‐dithiophenes with very efficient work‐up procedure is presented. The molecular structures of two silyl functionalized dibromo‐dithiophenes in the solid state have been determined to investigate the structural influences of different functional groups on the degree of π‐conjugation within the dithiophene moieties, as well as their packing properties. The planar alignment of the tert‐butyldimethylsilyl‐functionalized dibromo‐dithiophene shows a significantly higher degree of conjugation of the π‐system with a more favorable molecular packing than the skewed arrangement of the triisopropylsilyl‐substituted species. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

14.
Structural studies and morphological features of a new family of linear, aliphatic even–even, X 34‐nylons, with X = 2, 4, 6, 8, 10, and 12, are investigated with X‐ray diffraction and electron microscopy. Solution‐grown crystals were obtained by isothermal crystallization from N,N‐dimethylformamide solutions. The thickness of lamellar‐like crystals was orders of magnitude less than the chain lengths of the polymer samples used, implying that the chains fold to form chain‐folded lamellae. The results bear a close resemblance, with the noticeable exception of 2 34‐nylon, to those reported for nylon 6 6 and other even–even nylon chain‐folded lamellar crystals. The basic structure of the straight‐stem lamellar core is similar to that of the classic nylon 6 6 triclinic α structure, and the chains tilt ≈42° relative to the lamellar normal. In the case of 2 34‐nylon, the structure resembles the 2 Y nylon series, and the chain tilt angle reduces to 36.6°. These combined results suggest that, even with a relatively low frequency of amide units along the backbone of these molecules, hydrogen bonding is still the dominant element in controlling the behavior, structure, and properties of these polymers. In addition, gels were prepared in concentrated sulfuric acid, and gel‐spun fibers were studied using X‐ray diffraction. The data are interpreted in terms of a modified nylon triclinic α structure that bears a resemblance to the structure of even–even nylons at elevated temperatures. © 2002 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 40: 2685–2692, 2002  相似文献   

15.
o‐Formylphenylboronic acid reacts with morpholine to form 1,3‐dihydro‐1‐hydroxy‐3‐morpholin‐4‐yl‐2,1‐benzoxaborole. The typical hydrogen‐bonded dimer motif with a planar benzoxaborole fragment has been obtained in the solid state. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

16.
The reaction of [Cp*Ir(bzpy)NO3] ( 1 ; bzpy=2‐benzoylpyridine, Cp*=pentamethylcyclopentadienyl anion), a competent water‐oxidation catalyst, with several oxidants (H2O2, NaIO4, cerium ammonium nitrate (CAN)) was studied to intercept and characterize possible intermediates of the oxidative transformation. NMR spectroscopy and ESI‐MS techniques provided evidence for the formation of many species that all had the intact Ir–bzpy moiety and a gradually more oxidized Cp* ligand. Initially, an oxygen atom is trapped in between two carbon atoms of Cp* and iridium, which gives an oxygen–Ir coordinated epoxide, whereas the remaining three carbon atoms of Cp* are involved in a η3 interaction with iridium ( 2 a ). Formal addition of H2O to 2 a or H2O2 to 1 leads to 2 b , in which a double MeCOH functionalization of Cp* is present with one MeCOH engaged in an interaction with iridium. The structure of 2 b was unambiguously determined in the solid state and in solution by X‐ray single‐crystal diffractometry and advanced NMR spectroscopic techniques, respectively. Further oxidation led to the opening of Cp* and transformation of the diol into a diketone with one carbonyl coordinated at the metal ( 2 c ). A η3 interaction between the three non‐oxygenated carbons of “ex‐Cp*” and iridium is also present in both 2 b and 2 c . Isolated 2 b and mixtures of 2 a – c species were tested in water‐oxidation catalysis by using CAN as sacrificial oxidant. They showed substantially the same activity than 1 (turnover frequency values ranged from 9 to 14 min?1).  相似文献   

17.
The titanocene acetylene complex [Cp*2Ti(η2-Me3SiC≡CSiMe3)] ( 14 ) reacts with 1-alkynes such as phenylacetylene ( 15 a ), 1-hexyne ( 15 b ), 1-dodecyne ( 15 c ) and trimethylsilylacetylene ( 15 d ) by ligand exchange and proton shift, to yield exclusively the 1-alkenyltitanocene acetylides [Cp*2Ti(CH=CHR)(C≡CR)] ( 21 ) (R = Ph ( 21 a ), CH3(CH2)3 ( 21 b ), CH3(CH2)9 ( 21 c ), SiMe3 ( 21 d )). The X-ray structure of 21 a is presented. In reaction of acetylene HC≡CH ( 15 e ) with 14 other products are formed. However, no intermediates, like [Cp*2Ti(η2-RC≡CH)] ( 17 ), [Cp*2Ti(H)C≡CR] ( 17 ) or [Cp*2Ti=C=CHR] ( 22 ) in reactions of 14 with 15 are detectable. On the other hand, a stable titanocenehydride [Cp*2Ti(H)OCH3] ( 23 ) is obtained by oxidative addition of CH3OH with Cp*2Ti, generated from 14 .  相似文献   

18.
The first coordination compound of 1,4‐dihydro‐2,3‐quinoxalinedione in ketoamine tautomeric form (denoted as H2qdione) was reported. H2qdione was obtained by a solid‐state reaction of o‐phenylenediamine and oxalic acid. Reaction of this ligand with CdCl2 solvothermally yielded a coordination polymer [Cd(H2qdione)Cl2]n, which was structurally characterized by X‐ray diffraction and IR spectroscopy. Continuous Cd2Cl2 diamonds form a double‐sided comb with terminal H2qdione‐κ2O,O′ as the comb teeth. Interaction of these combs through very extensive π–π stacking, C–H ··· Cl, and N–H ··· Cl hydrogen bonds leads to a novel 3D architecture and significant enhancement of solid‐state luminescence of about 10 times compared to the free H2qdione ligand.  相似文献   

19.
The isotypic nitridosilicates Li4Ca3Si2N6 and Li4Sr3Si2N6 were synthesized by reaction of strontium or calcium with Si(NH)2 and additional excess of Li3N in weld shut tantalum ampoules. The crystal structure, which has been solved by single‐crystal X‐ray diffraction (Li4Sr3Si2N6: C2/m, Z = 2, a = 6.1268(12), b = 9.6866(19), c = 6.2200(12) Å, β = 90.24(3)°, wR2 = 0.0903) is made up from isolated [Si2N6]10– ions and is isotypic to Li4Sr3Ge2N6. The bonding angels and distances within the edge‐sharing [Si2N6]10– double‐tetrahedra are strongly dependent on the lewis acidity of the counterions. This finding is discussed in relation to the compounds Ca5Si2N6 and Ba5Si2N6, which also exhibit isolated [Si2N6]10– ions.  相似文献   

20.
Deep blue‐violet colored powder samples of Ag2ZnZr2F14 were synthesized by heating Zn(NO3)2·4H2O, Ag and ZrOCl2·8H2O at 300 °C under fluorine atmosphere. The crystal structure of Ag2ZnZr2F14 was refined from X‐ray powder diffraction data using the Rietveld method (C2/m, a = 9.0206(1) Å, b = 6.6373(1) Å, c = 9.0563(1) Å, β = 90.44(1)°, Z = 2). The structure is derived from the isotypic Ag3Zr2F14 by replacing only one of the two crystallographically different Ag2+ ions with Zn2+ ions, thus leading to discrete Ag2F7 dimers. These dimers are connected via nearly linear Ag–F···F–Ag bridges with short F···F distances of 2.33 Å to form two‐legged ladders. Magnetic susceptibility measurements and density functional calculations show that the two Ag2+ ions in each Ag2F7 dimer are strongly coupled antiferromagnetically.  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