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 62 毫秒
1.
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   

2.
Abstract

This review covers 30?Pt(II) complexes in which the inner coordination of PtP2S2 is created by organodiphosphines and a pair of S-donor ligands. The organodiphosphines generate four-(PCP), five-(PC2P), six-(PC3P) and seven-(PC4P) membered metallocyclic rings. These complexes crystallize in three crystal systems: orthorhombic (four examples), triclinic (nine examples) and monoclinic (seventeen examples). The respective metallocyclic rings open in the sequence: (mean values): 73.2° (PCP) < 86.0° (PC2P) < 94.5° (PC3P) < 97.5° (PC4P). The mean values of Pt-P and Pt-S bond distances are 2.247 and 2.350?Å, respectively. The structure parameters are analyzed and discussed with attention to any trans-influence.  相似文献   

3.
The photochemical carbonyl substitution of (η6-anisole)Cr(CO)3 has been investigated by laser flash photolysis. Both transient spectra and second-order rate constants for the reactions of transients with nucleophiles are found to be extremely variable depending upon solvents used. The coordination of benzene to the transient in cyclohexane forms the transient in benzene, indicating two discrete chemical species: (η6-anisole)Cr(CO)2 and (η6-anisole)Cr(CO)22-benzene). The latter type of transient was observed also for fluorobenzene and mesitylene, leading to the assignment of a weak band in the visible region as η2-arene → Cr charge transfer. The existence of (η6-arene)Cr(CO)22-arene′) may throw light on what have been described as solvent effects in organometallic reactions.  相似文献   

4.
5.
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   

6.
7.
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   

8.
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   

9.
Diuranium μ-η(6),η(6)-arene complexes supported by ketimide ligands were synthesized and characterized. Disodium or dipotassium salts of the formula M(2)(μ-η(6),η(6)-arene)[U(NC(t)BuMes)(3)](2) (M = Na or K, Mes = 2,4,6-C(6)H(2)Me(3)) and monopotassium salts of the formula K(μ-η(6),η(6)-arene)[U(NC(t)BuMes)(3)](2) (arene = naphthalene, biphenyl, trans-stilbene, or p-terphenyl) were both observed. Two different salts of the monoanionic, toluene-bridged complexes are also described. Density functional theory calculations have been employed to illuminate the electronic structure of the μ-η(6),η(6)-arene diuranium complexes and to facilitate the comparison with related transition-metal systems, in particular (μ-η(6),η(6)-C(6)H(6))[VCp](2). It was found that the μ-η(6),η(6)-arene diuranium complexes were isolobal with (μ-η(6),η(6)-C(6)H(6))[VCp](2) and that the principal arene-binding interaction was a pair of δ bonds (total of 4e) involving both metals and the arene lowest unoccupied molecular orbital. Reactivity studies have been carried out with the mono- and dianionic μ-η(6),η(6)-arene diuranium complexes, revealing contrasting modes of redox chemistry as a function of the system's state of charge.  相似文献   

10.
《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   

11.
A series of mononuclear lanthanide(III) complexes [Ln(LH(2))(H(2)O)(3)Cl](ClO(4))(2) (Ln = La, Nd, Sm, Eu, Gd, Tb, Lu) of the tetraiminodiphenolate macrocyclic ligand (LH(2)) in 95 : 5 (v/v) methanol-water solution fix atmospheric carbon dioxide to produce the carbonato-bridged trinuclear complexes [{Ln(LH(2))(H(2)O)Cl}(3)(μ(3)-CO(3))](ClO(4))(4)·nH(2)O. Under similar conditions, the mononuclear Y(III) complex forms the dimeric compound [{Y(LH(2))(H(2)O)Cl}(μ(2)-CO(3)){Y(LH(2))(H(2)O)(2)}](ClO(4))(3)·4H(2)O. These complexes have been characterized by their IR and NMR ((1)H, (13)C) spectra. The X-ray crystal structures have been determined for the trinuclear carbonato-bridged compounds of Nd(III), Gd(III) and Tb(III) and the dinuclear compound of Y(III). In all cases, each of the metal centers are 8-coordinate involving two imine nitrogens and two phenolate oxygens of the macrocyclic ligand (LH(2)) whose two other imines are protonated and intramolecularly hydrogen-bonded with the phenolate oxygens. The oxygen atoms of the carbonate anion in the trinuclear complexes are bonded to the metal ions in tris-bidentate μ(3)-η(2):η(2):η(2) fashion, while they are in bis-bidentate μ(2)-η(2):η(2) mode in the Y(III) complex. The magnetic properties of the Gd(III) complex have been studied over the temperature range 2 to 300 K and the magnetic susceptibility data indicate a very weak antiferromagnetic exchange interaction (J = -0.042 cm(-1)) between the Gd(III) centers (S = 7/2) in the metal triangle through the carbonate bridge. The luminescence spectral behaviors of the complexes of Sm(III), Eu(III), and Tb(III) have been studied. The ligand LH(2) acts as a sensitizer for the metal ions in an acetonitrile-toluene glassy matrix (at 77 K) and luminescence intensities of the complexes decrease in the order Eu(3+) > Sm(3+) > Tb(3+).  相似文献   

12.
DFT calculations were carried out for the Ni(0) complex Ni(PH(3))(2)(η(2)-CO(2)), which is a model compound for the well-known Ni(0) carbon dioxide complexes containing various tertiary phosphane ligands. The electronic structure of the complex was elucidated using domain-averaged Fermi hole (DAFH), quantum theory of atoms in molecules (QTAIM), electron localization function (ELF), charge decomposition analysis (CDA), and natural bond orbital (NBO) methods. The carbon dioxide ligand in the complex reveals an unexpected coordination behavior. Apart from the expected π-donation interaction, the C-O σ bond takes also part in the electron donation. Moreover, the back-donation is slightly influenced by the phosphorus atom adjacent to the noncoordinated O of carbon dioxide as it transfers electron density directly to carbon. This unconventional way of back-donation may also explain the bent character of the Ni-C bond path. Due to excess kinetic energy density, no bond critical point was found between the coordinating oxygen and the nickel center. A strong relationship has been found between the DAFH and the NBO methods, which can provide additional information for the interpretation of DAFH eigenvectors.  相似文献   

13.
To get further insight in the reaction of nucleophilic substitution upon changing the ligand trans to a η(2)-olefin, the reactivity of some monoanionic platinum(II) complexes (trans-[PtCl(2)X(η(2)-C(2)H(4))](-), X = Cl(-), 1, OH(-), 2, and CH(2)NO(2)(-), 3) towards pyridines with different steric hindrance (py, 4-Mepy, and 2,6-Me(2)py) has been tested. All crystallographic (2 and 3 reported for the first time) and spectroscopic data are in accord with a platinum-olefin interaction decreasing in the order 2 > 1 > 3, paralleling the decreasing electronegativity of the donor atom (O > Cl > C). Not only the platinum-olefin bond but also the bond between platinum and the ligand trans to the olefin appear to be strongest in 2 (Pt-O distance at the lower limit for this type of bond). In the reaction with py, the ligand trans to the olefin is displaced in 1 and 2. Moreover the reaction is in equilibrium in the case of sterically hindered 2,6-Me(2)py, the equilibrium being shifted moderately or prevalently toward the reagents in the case of 1 and 2, respectively. In the case of 3, the reaction with pyridines leads to substitution of the olefin instead of the carbanion. This is in accord with the observation that carbanions strongly weaken the trans Pt-olefin bond.  相似文献   

14.
《Polyhedron》1987,6(2):261-267
A new, high-yield method has been developed for the preparation of MO(CO)2I22-dppm)(η1-dppm). The title compound was prepared by the reaction of [Et4N][Mo(CO)4I3] with dppm in benzene in 95% yield. It has been characterized by a single-crystal X-ray study. The crystallographic data are as follows: monoclinic, space group P21/n, a = 19.023(4) Å, b = 14.439(3) Å, c = 20.141(5) Å, β = 100.45(2)°, V = 5440(2) Å3 Z = 4. The geometry around the central metal atom could be considered as either a distortion from a capped octahedron with a carbonyl in a capping position or from a trigonal prism with the iodine capping a rectangular face. The solution behavior of Mo(CO)2I2(dppm)2 was examined with 31P NMR, which showed it to be fluxional.  相似文献   

15.
INTRODUCTION

As discussed in a recent preliminary publication1, the complex [Ru(η4-C8H12){η-P3C2But 2CH(SiMe3)2}] (1) (C8H12 = cycloocta-l,5-diene) was prepared by the reaction of [Ru(η6-C10H8)(η4-C8H12)] (2) (C10H8 = naphthalene) with the 1,2,4-triphosphole P3C2But 2CH(SiMe3)2 (3) (Fig. 1), illustrating the aromatic behaviour of (3).  相似文献   

16.
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   

17.
Visible light irradiation of the dicarbollide complex [(η-9-SMe2-7,8-C2B9H10)Fe(η-C6H6)]+ (2a) in the presence of the benzene derivatives in CH2Cl2/MeNO2 resulted in cations [(η-9-SMe2-7,8-C2B9H10)Fe(η-C6R6)]+ (2b-g; arene is anisole (b), toluene (c), m-xylene (d), mesitylene (e), durene (f), and hexamethylbenzene (g)) due to the arene exchange. The structures of [2g]PF6 and related tricarbollide complex [(η-1-ButNH-1,7,9-C3B8H10)Fe-(η-C6H6)]PF6 (1) were confirmed by X-ray diffraction analysis. The nature of bonding in cations 1, 2a, and [CpFe(η-C6H6)]+ was analyzed by an energy decomposition analysis.  相似文献   

18.
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   

19.
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   

20.
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   

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