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 538 毫秒
1.
Ferrocenyl substituted ruthenium metallacyclic compounds, [Ru2(CO)6{μ-η1122-1,4-Fc2C5H2O}] (1) and [Ru2(CO)6{μ-η1122-1,5-Fc2C5H2O}] (2) have been synthesized and structurally characterized. Electrochemical studies for 1 and 2 and the respective quinone derivatives 3 and 4 show weak to no electrochemical coupling at the mixed-valent intermediate state which is dependent on the complex frameworks.  相似文献   

2.
The lithium complex with the acenaphthylene dianion [Li(Et2O)2]22:3[Li(3:3-C12H8)]2 (1) was synthesized by the reduction of acenaphthylene with lithium in diethyl ether. According to the X-ray diffraction data, compound 1 has a reverse-sandwich structure with the bridging dianion 2:3[Li(3:3-C12H8)]2. Two lithium atoms in complex 1 are located between two coplanar acenaphthylene ligands of the 2:3[Li(3:3-C12H8)]2 2– dianion and are 3-coordinated with the five- and six-membered rings. The lanthanum complex with the acenaphthylene dianion [LaI2(THF)3]2(2-C12H8) (2) was synthesized by the reduction of acenaphthylene in THF with the lanthanum(iii) complex [LaI2(THF)3]2(2-C10H8) containing the naphthalene dianion. The 1H NMR spectrum of complex 2 in THF-d8 exhibits four signals of the acenaphthylene dianion, whose strong upfield shifts compared to those of free acenaphthylene indicate the dianionic character of the ligand. The highest upfield chemical shift belongs to the proton bound to the C atom on which, according to calculation, the maximum negative charge is concentrated.  相似文献   

3.
The first μ-η(2):η(2)-diselenidodicopper(II) complex has been obtained in the reaction of a copper(I) complex with N,N',N″-tribenzyl-cis,cis-1,3,5-triaminocyclohexane and elemental selenium. The structure and reactivity of the complex is described.  相似文献   

4.
Reaction of the cluster Os3(μ-CO)(CO)93112-Me3SiC2Me) with HC≡CCOOMe in benzene at 70 °C results in Os3(CO)931122-C(SiMe3)C(Me)C(COOMe)CH× (5), Os3(CO)931122-C(SiMe3)C(Me)C(H)C(COOMe)CH× (6), Os3(CO)9{μ-η114-C(SiMe3)C(Me)C(H)C(COOMe)CH× (7), and Os3(CO)δ31141-C(SiMe3)C(Me)C(H)C(COOMe)× complexes (8), containing an osmacyclopentadiene moiety. Complexes5–8 were characterized by1H NMR and IR spectroscopy. The structure of clusters5 and8 was confirmed by X-ray analysis. Complex7 is formed from cluster5 as a result of a new intramolecular rearrangement and complex8 is obtained by decarbonylation of compound6. Complex8 adds PPh3 to give Os3(CO)δ(PPh3){μ-η114-C(SiMe3)C(Me)C(H)C(COOMe)×.  相似文献   

5.
6.
1 INTRODUCTION Since last decade, the chemistry of divalent organolanthanide complexes has yielded particularly remarkable and striking results. The major break- through in the chemistry of divalent organolan- thanides, especially Sm(II), includes the u…  相似文献   

7.
Abstract

In this review structural parameters of forty complexes with an inner coordination sphere of Pt(η2-P2L)(η2-S2L) are analyzed and classified These complexes crystallize in three crystal systems: orthorhombic (four examples), triclinic (six examples) and monoclinic (thirty examples). The organodiphosphines create four- (PCP), five- (PC2P), six- (PC3P) and seven- (PC4P) membered metallocyclic rings with mean P-Pt-P bite angle values of 72.5° (PCP) < 85.3° (PC2P) < 93.0° (PC3P) < 97.4° (PC4P). The dithiolates create four- (SCS), five- (SC2S), six- (SC3S; SCSCS; SPNPS; SPCPS) and seven- (SC4S) membered metallocyclic rings with mean S-Pt-S bite angle values of 74.5° (SCS) < 85.8° (SCSCS) < 87.0° (SPNPS) < 89.0° (SC2S) < 92.3° (SC4S) < 93.5° (SC3S) < 97.5° (SPCPS). The mean Pt-P and Pt-S bond distances are 2.257 and 2.328?Å, respectively. The data are compared with those found in complexes with inner coordination spheres of Pt(PL)2(SL)2, Pt(PL)22-S2L) and Pt(η2-P2L)(SL)2.  相似文献   

8.
The electron distributions and bonding in Ru3(CO)9( 3- 2, 2, 2-C6H6) and Ru3(CO)9( 3- 2, 2, 2-C60) are examined via electronic structure calculations in order to compare the nature of ligation of benzene and buckminsterfullerene to the common Ru3(CO)9 inorganic cluster. A fragment orbital approach, which is aided by the relatively high symmetry that these molecules possess, reveals important features of the electronic structures of these two systems. Reported crystal structures show that both benzene and C60 are geometrically distorted when bound to the metal cluster fragment, and our ab initio calculations indicate that the energies of these distortions are similar. The experimental Ru–Cfullerene bond lengths are shorter than the corresponding Ru–Cbenzene distances and the Ru–Ru bond lengths are longer in the fullerene-bound cluster than for the benzene-ligated cluster. Also, the carbonyl stretching frequencies are slightly higher for Ru3(CO)9( 3- 2, 2, 2-C60) than for Ru3(CO)9( 3- 2, 2, 2-C6H6). As a whole, these observations suggest that electron density is being pulled away from the metal centers and CO ligands to form stronger Ru–Cfullerene than Ru–Cbenzene bonds. Fenske-Hall molecular orbital calculations show that an important interaction is donation of electron density in the metal–metal bonds to empty orbitals of C60 and C6H6. Bonds to the metal cluster that result from this interaction are the second highest occupied orbitals of both systems. A larger amount of density is donated to C60 than to C6H6, thus accounting for the longer metal–metal bonds in the fullerene-bound cluster. The principal metal–arene bonding modes are the same in both systems, but the more band-like electronic structure of the fullerene (i.e., the greater number density of donor and acceptor orbitals in a given energy region) as compared to C6H6 permits a greater degree of electron flow and stronger bonding between the Ru3(CO)9 and C60 fragments. Of significance to the reduction chemistry of M3(CO)9( 3- 2, 2, 2-C60) molecules, the HOMO is largely localized on the metal–carbonyl fragment and the LUMO is largely localized on the C60 portion of the molecule. The localized C60 character of the LUMO is consistent with the similarity of the first two reductions of this class of molecules to the first two reductions of free C60. The set of orbitals above the LUMO shows partial delocalization (in an antibonding sense) to the metal fragment, thus accounting for the relative ease of the third reduction of this class of molecules compared to the third reduction of free C60.  相似文献   

9.
Liu X  Wang X  Wang Q  Andrews L 《Inorganic chemistry》2012,51(13):7415-7424
Infrared spectra of the matrix isolated OMS, OM(η(2)-SO), and OM(η(2)-SO)(η(2)-SO(2)) (M = Ti, Zr, Hf) molecules were observed following laser-ablated metal atom reactions with SO(2) during condensation in solid argon and neon. The assignments for the major vibrational modes were confirmed by appropriate S(18)O(2) and (34)SO(2) isotopic shifts, and density functional vibrational frequency calculations (B3LYP and BPW91). Bonding in the initial OM(η(2)-SO) reaction products and in the OM(η(2)-SO)(η(2)-SO(2)) adduct molecules with unusual chiral structures is discussed.  相似文献   

10.
11.
The polymerization of styrene, methyl methacrylate, and vinyl chloride catalyzed by η5-cyclopentadienyl-η2-styrenedicarbonylmanganese is studied. It is shown that the cyclopentadienyl complex of manganese containing the monomer ligand (styrene) in the coordination sphere can initiate the radical polymerization of vinyl monomers in a mild temperature range. On the basis of the experimental data and the quantum-chemical simulation of the initial stages of the process, schemes describing the initiation of polymerization under the action of the complex under study and the binary initiating system containing carbon tetrachloride are advanced. In the latter case, additional acceleration of the reaction is related to the interaction of carbon tetrachloride with the triplet form of the manganese complex that yields trichloromethyl radicals initiating polymerization.  相似文献   

12.
We report the oxidative dinuclear addition of a Pd(I)-Pd(I) bond to arenes. The oxidative dinuclear addition products, which have a bi-π-allyl-type arene dipalladium(II) structure, were obtained from [2.2]paracyclophane, anthracene, tetracene, and pentacene. A systematic study of the reaction of [Pd(2)(CH(3)CN)(6)][BF(4)](2) with benzene and polyacenes showed that the larger polyacenes, tetracene and pentacene, afforded the oxidative dinuclear addition products, while benzene, naphthalene, and anthracene gave the π-sandwich Pd(I)-Pd(I) complexes.  相似文献   

13.
New hexadentate dinucleating ligands having a xylyl linker, X–L–R, were synthesized, where X–L–R = 1,3-bis[bis(6-methyl-2-pyridylmethyl)aminomethyl]-2,4,6-trimethybenzene (Me2–L–Me) and 1,3-bis[bis(6-methyl-2-pyridylmethyl)aminomethyl]-2-fluorobenzene (H–L–F). They form dinuclear copper(I) complexes, [Cu2(X–L–R)]2+ (Me2–L–Me (1) and H–L–F (2)). The copper(I) complexes in acetone at −78 °C react with O2 to produce intra- and intermolecular (μ-η22-peroxo)dicopper(II) species depending on the concentrations of the complexes:  both complexes generate intramolecular (μ-η22-peroxo)dicopper(II) species [Cu2(O2)(X–L–R)]2+ (1-O2 and 2-O2) at the concentrations below ∼5 mM, whereas at ∼60 mM, both complexes produce intermolecular (μ-η22-peroxo)dicopper(II) species, which were confirmed by the electronic and resonance Raman spectroscopies.  The electronic spectrum of 1-O2 in acetone at concentrations below ∼5 mM showed an absorption band at (λmax = 442 nm, ε = 5600 M−1 cm−1) assignable to the πσ1(O–O)-to-Cu(II) ((dx2-y2+dx2-y2) component) LMCT transition in addition to an intense band attributable to the πσ1(O–O)-to-Cu(II) ((dx2-y2-dx2-y2) component) LMCT transition (λmax = 359 nm, ε = 21000 M−1 cm−1), indicating that the (μ-η22-peroxo)Cu(II)2 core of 1-O2 takes a butterfly structure. Decomposition of 1-O2 resulted in hydroxylation of the 2-position of the xylyl linker with 1,2-methyl migration (NIH shift), suggesting that the hydroxylation reaction proceeds via a cationic intermediate as proposed for closely related (μ-η22-peroxo)Cu(II)2 complexes having a xylyl linker. Kinetic study of the decomposition (hydroxylation of the xylyl linker) of 1-O2 suggests that a stereochemical effect of the methyl group in the 2-position of the xylyl linker has a significant influence on a transition state for decomposition (hydroxylation of the xylyl linker).  相似文献   

14.
《Polyhedron》2001,20(9-10):1065-1070
Decomplexation of Ca3(thd)6 by mono- and bidentate N-donors [morpholine, dimorpholinoethane (DIMOE), TMEDA, bipyridine] afforded the corresponding adducts Ca(thd)2L [L=morpholine (1a), DIMOE (1b), TMEDA (2)] and {Ca(thd)2}2(bipy) (3). All complexes have been fully characterised by elemental analysis, FT-IR and 1H NMR spectroscopy. Compounds 1b and 3 have also been characterised by X-ray crystallography. The structure of 3 is based on six- and seven-coordinated Ca centres; it is the first dimeric volatile Lewis base adduct of Ca(thd)2. The thermal behaviour of all derivatives has been studied by thermal gravimetric analysis.  相似文献   

15.
The synthesis, electronic structure, and reactivity of a uranium metallacyclopropene were comprehensively studied. Addition of diphenylacetylene (PhC≡CPh) to the uranium phosphinidene metallocene [η5-1,2,4-(Me3C)3C5H2]2U=P-2,4,6-tBu3C6H2 ( 1 ) yields the stable uranium metallacyclopropene, [η5-1,2,4-(Me3C)3C5H2]2U[η2-C2Ph2] ( 2 ). Based on density functional theory (DFT) results the 5f orbital contributions to the bonding within the metallacyclopropene U-(η2-C=C) moiety increases significantly compared to the related ThIV compound [η5-1,2,4-(Me3C)3C5H2]2Th[η2-C2Ph2], which also results in more covalent bonds between the [η5-1,2,4-(Me3C)3C5H2]2U2+ and [η2-C2Ph2]2− fragments. Although the thorium and uranium complexes are structurally closely related, different reaction patterns are therefore observed. For example, 2 reacts as a masked synthon for the low-valent uranium(II) metallocene [η5-1,2,4-(Me3C)3C5H2]2UII when reacted with Ph2E2 (E=S, Se), alkynes and a variety of hetero-unsaturated molecules such as imines, ketazine, bipy, nitriles, organic azides, and azo derivatives. In contrast, five-membered metallaheterocycles are accessible when 2 is treated with isothiocyanate, aldehydes, and ketones.  相似文献   

16.
Laser-ablated ruthenium atoms undergo reaction with acetylene during condensation in excess neon and argon matrices to form a metallacycle complex, insertion into the C-H bond, and rearrangement to the vinylidene complex. The subject molecules were identified by (13)C(2)H(2) and C(2)D(2), isotopic substitutions and density functional theory (DFT) frequency calculations. The HRuCCH molecule is described by Ru-H, CH, and CC stretching modes and CCH deformation modes. A very strong CC double bond stretching, weak CH stretching, and CCH deformation frequencies were observed for the Ru═C═CH(2) complex. The metallacycle Ru-η(2)-(C(2)H(2)) is characterized through CC double bond stretching, CH stretching and CCH deformation modes. The reaction mechanism for formation of the Ru═C═CH(2) complex was investigated by B3LYP internal reaction coordinate calculations, and the hydrido-alkyny complex is the rate-determining step. The delocalized three-center-four-electron π bond using the Ru 4d(xz) electron pair contributes to the C-C π* orbital and provides stabilization energy (ΔE((2)), second-order perturbation) for the vinylidene Ru═C═CH(2) complex.  相似文献   

17.
18.
The oxidation of [(Cp*Mo)2(μ,η66-P6)] ( 1 ) with halogens or halogen sources was investigated. The iodination afforded the ionic complexes [(Cp*Mo)2(μ,η33-P3)(μ,η1111-P3I3)][X] (X=I3, I) ( 2 ) and [(Cp*Mo)2(μ,η44-P4)(μ-PI2)][I3] ( 3 ), while the reaction with PBr5 led to the complexes [(Cp*Mo)2(μ,η33-P3)(μ-Br)2][Cp*MoBr4] ( 4 ) [(Cp*MoBr)2(μ,η33-P3)(μ,η1-P2Br3)] ( 5 ) and [(Cp*Mo)2(μ-PBr2)(μ-PHBr)(μ-Br)2] ( 6 ). The reaction of 1 with the far stronger oxidizing agent PCl5 was followed via time- and temperature-dependent 31P{1H} NMR spectroscopy. One of the first intermediates detected at 193 K was [(Cp*Mo)2(μ,η33-P3)(μ-PCl2)2][PCl6] ( 8 ) which rearranges upon warming to [(Cp*Mo)2(μ-PCl2)2(μ-Cl)2] ( 9 ), [(Cp*MoCl)2(μ,η33-P3)(μ-PCl2)] ( 10 ) and [(Cp*Mo)2(μ,η44-P4)(μ-PCl2)][Cp*MoCl4] ( 11 ), which could be isolated at room temperature. All complexes were characterized by single-crystal X-ray diffraction, NMR spectroscopy and their electronic structures were elucidated by DFT calculations.  相似文献   

19.
The complex [(η(7)-C(7)H(7))Zr{N(SiMe(3))(2)}(thf)] (2) loses THF upon sublimation to afford a chain polymer consisting of [(η(7)-C(7)H(7))Zr{N(SiMe(3))(2)}] (3) units; they are connected by cycloheptatrienyl ligands in an unprecedented antifacial η(7):η(2)-bridging mode. The basicity of the bis(trimethylsilyl)amido ligand can also be used to introduce other ligands by acid-base reactions.  相似文献   

20.
Treatment of [W(CO)(MeC2Me)2(-C5H5)][PF6] with ONMe3 in acetonitrile yields [W(NCMe)(MeC2Me)2(-C5H5)][PF6] which undergoes irreversible reduction at a Pt electrode in THF. Sodium amalgam reduction of [W(NCMe) (MeC2Me)2(-C5H5)][PF6] gives orange crystals of [W2(µ-,, 2, 2-C4Me4)2 (-C5H5)2] X-ray studies on which reveal pairwise alkyne coupling and a novel bis(metallacyclopentadiene) structure.Dedicated to Professor L. F. Dahl on the occasion of his 65th birthday.  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