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.
The potential of several alkylcobalt complexes as catalysts for hydrogenation and isomerization of alkenes has been investigated. The complexes CH3Co(CO)2(Pom-Pom) (Pom-Pom = 1,2 bis(dimethoxyphosphino)ethane), CH3Co(CO)3P(OMe)3 and C6H5CH2Co(CO)3PPh 3 are compared to CH3Co(CO)2(P(OMe)3)2, for their ability to function in catalytic cycles. Each is active for hydrogenation and isomerization of alkenes under conditions where the carbonylation-decarbonylation equilibrium is readily established. The lifetime for the complexes is much shorter than for CH3Co(CO)2 (P(OMe)3)2 suggesting that two phosphorus donors in trans positions in an intermediate is a requirement for catalyst stability in these alkylcobalt complexes.  相似文献   

2.
The Stoichiometry of thermal decomposition was studied for the following compounds: Ni(NCS)2(2-Mepy)2 (I), (Me=methyl, py=pyridine), Ni(NCS)2(2-Etpy)2 (II) (Et=ethyl), Ni(NCS)2(2-Clpy)2 (III), Ni(NCS)2(2-Brpy)2 (IV), Ni(NCS)2(2-NH2py)2 (V), Ni((NCS)2(2-NH2py)2·3/4 (C2H5)2O (VI). The release of volatile ligands 2-Rpy is a one-step process for complexes I, II, III and IV, while for V and VI it is a two-step process, Ni(NCS)2(2-NH2py)1 (VII) being formed as an intermediate complex. It was found that complexes I and II are square-planar; the others exhibited pseudo-octahedral geometry. The differences in stereochemistry of the above complexes are explained by the different electronic properties of 2-Rpy.  相似文献   

3.
Protonation of the dinuclear compounds [M2(μ-CO)(CO)4(μ-R2PYPR2)2] by HBF4 or HPF6 leads to the formation of crystalline cationic hydrido products [M2H(CO)5(μ-R2PYPR2)2]X and [M2(μ-H)(μ-CO)(CO)4(μ-R2PYPR2)2]X (X = BF4 or PF6) in which the hydride ligand is terminal for M = Ru, Y = N(Et) and R = OMe or OPri and bridging for M = Fe, Y = CH2 and R = Me or Ph, for M = Fe, Y = N(Et) and R = OMe, OEt, OPri or OPh and for M = Ru, Y = CH2 and R = Ph; the fluxional behaviour of [Ru2H(CO)5{μ-(RO)2PN(Et)P(OR)2}2]+ (R = Me or Pri) in solution is described.  相似文献   

4.
The isolation of non-fluxional alumoxane compounds, [(tBu)2Al{OAl(tBu)2}]2 and [(tBu)AlO]n (n = 6, 7, 8, 9), has allowed for an investigation of the mode of activity observed for alumoxanes as co-catalysts for the zirconocene polymerization of olefins. [(tBu)2Al{OAl(tBu)2}]2, which contains two three-coordinate aluminum centers, shows no reaction with Cp2ZrMe2, and no catalytic activity towards ethylene polymerization. In contrast, the closed cage compound [(tBu)AlO]6 reacts reversibly to give the ion pair complex, [Cp2ZrMe][(rBu)6Al6O6Me], which is active as a catalyst for the polymerization of ethylene. Polymerization is also observed for mixtures of Cp2ZrMe2 with [(tBu)AlO]n (n = 7, 9). A new concept, “latent Lewis acidity”, is proposed to account for the reactivity of the cage alumoxanes, [(tBu)AlO]n.  相似文献   

5.
The kinetics and mechanisms of propadiene polymerization under the influence of [Rh(CO)2Cl]2, Rh(CO)2P(C6H5)3Cl, Rh(CO)3Cl are reported. The reaction rates are first-order in Rh(CO)2P(C6H5)3Cl and Rh(CO)3Cl and half-order in [Rh(CO)2Cl]2. They are second-order in the substrate for Rh(CO)3Cl and [Rh(CO)2Cl]2 and first-order for Rh(CO)2P(C6H5)3Cl. The data are interpreted in terms of a common intermediate mechanism. The formation of this common intermediate is the rate-determining step. A solvent effect is also discussed.  相似文献   

6.
The hydrolysis of aryltellurium trihalides to aryltellurium oxide halides in neutral aqueous media is considered to proceed in a stepwise manner in which the first stage involves the formation of a monomeric species. In alkaline media the initially isolated hydrolysis product analyses as (p-EtOC6H4)TeO(OH) and on treatment with dilute acid, affords the known (p-EtOC6H4TeO)2O.The infrared spectra are assigned in the low frequency region for the compounds RTe(O)X (X = halogen), RTeO(OH) and (RTeO)2O. It is argued that the probable coordination number for tellurium in RTe(O)X is four, and that a ring structure is likely. The preparation of the salts(C5H5NH+)(RTeCl4?) is reported and the anions are considered to have square-based pyramidal structures of approximately C4v symmetry. The reaction of (C5H5NH+)(PhTeCl4?) with acetone affords Ph(CH3COCH2)TeCl2.  相似文献   

7.
Dicobalt octacarbonyl and some of its derivatives (NaCo(CO)4, Co4(CO)12, Hg[Co(CO)4]2, [Co(CO)3PPh3]2, NaCo(CO)3PPh3) react with activated gem-dihalides, R2CX2, such as dichlorodiphenylmethane, 9,9-dihalofluorenes and dimethyl dibromomalonate, to give the ‘dimer’ olefin, R2CCR2. The course of this conversion involves formation of the coupling product, R2XCCXR2, followed by dehalogenation of the latter. These separate steps have been confirmed for activated monohalides (bromodiphenylmethane, 9-bromofluorene, dimethyl bromomalonate) which were readily coupled by cobalt carbonyls, and for activated vicinal dihalides (D,L and meso-dibromostilbene, 9,9′-dichlorobisfluorenyl) which cobalt carbonyls readily dehalogenated. A radical mechanism is favored for these processes, and indirect evidence in its favor is presented.  相似文献   

8.
Crystal-structure line profile refinements have been made for neutron powder diffraction data collected at room temperature from the ternary uranium-thorium dicarbide U0.1Th0.9C2 and the two binary dicarbides UC2 and ThC2. U0.1Th0.9C2 is monoclinic, with unit cell dimensionsa = 6.630(2), b = 4.183(1), c = 6.690(2)A˚, β = 103.86(1)°, isostructural with ThC2, a = 6.684(2), b = 4.220(1), c = 6.735(2)A˚, β = 103.91(1)°. The tetragonal structure for UC2 is confirmed, witha = 3.522(1), c = 5.988(1)A˚. U0.1Th0.9C2 contains discrete C-C groups 1.297(8)A˚long; more precise values for C-C lengths have been determined for UC2 (1.322(4)A˚) and ThC2 (1.304(6)A˚).  相似文献   

9.
The aminolysis reactions of O-ethyl S-(Z-phenyl) dithiocarbonates (Z=p-CH3, H, p-Cl, and p-NO2) with anilines (AN) and N,N-dimethylanilines (DMA) in acetonitrile at 30.0°C are investigated. Relatively small values of βXnuc,0.4 ca. 0.7) and βZlg −0.1 ca. −0.4) for both ANs and DMAs, significantly large kH/kD values (1.1 ca. 1.9) involving deuterated anilines, and large negative ρXZ values for ANs (−0.56) are interpreted to indicate a concerted mechanism for both ANs and DMAs but with a hydrogen bonded four-center type transition state (TS) for ANs. The relative leaving ability, k(Z=p-NO2)/k(Z=p-CH3), is smaller for ANs than for DMAs, especially for a weaker nucleophile (1.9 and 4.7 for AN and DMA, respectively, with X=p-Cl). This suggests that the rate enhancement by the hydrogen-bond formation in the four-center type TS for AN is greater for a weaker nucleofuge (Z=p-CH3), especially when the nucleophile (X=p-Cl) is weaker. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet: 30: 419–423,1998  相似文献   

10.
Thienylmercury(II)chloride reacts with [Pd(PPh3)2Cl2], [Pd(PPh3)4] and [Pt(PPh3)4] to afford new compounds containing a metal-2-thienyl linkage. The compound [Pd(PPh3)2(2-C4H3S)Cl] probably has trans stereochemistry.2-Bromothiophen undergoes oxidative addition with [Pd(PPh3)4] and [Pt(PPh3)4], probably via a radical mechanism. With [Pd(CO)(PPh3)3], a carbonyl inserted product is obtained. The bromo-metal(II) complexes have trans stereochemistry. The course of the reaction between 3-methyl-2-bromothiophen and Pd(PPh3)4 is more complex. Thus, there is evidence of some cis bromopalladium(II) compounds amongst the products, also there is good evidence to support the view that some isomerisation of 3-methyl-2-thienyl to 4-methyl-2-thienyl occurs during the reaction, thus giving greater molar quantities of [Pd(PPh3)2(4-CH3-2-C4H2S)Br] than can be accounted for from any initial 4-methyl-2-bromothiophen impurity.The metallation of the thiophen ring, probably in the 4-position, with palladium(II) is described for 3-theylidene-4-methylaniline.  相似文献   

11.
Summary The ammoniation ofcis-[Rh(en)2Cl2] · (ClO4) in liquid NH3 was studied at constant ionic medium of 0.20 m perchlorate in the 0 to 35° range. The complex reacts in two distinct steps to givecis-[Rh(en)2(NH3)2] · (ClO4)3, with the intermediate formation ofcis-[Rh(en)2(NH3)Cl] · (ClO4)2. Both steps follow a conjugate-base mechanism. Activation parameters were obtained for the acid-base preequilibrium and the rate-determining step. The entropies of activation for the rate-determining step are 0 and –42 JK–1mol–1 for the first and second ammoniations respectively. These values are considerably lower than those found for the cobalt(III) analogues. The entropy changes for the acid-base equilibria are –84 and –36 JK–1mol–1 respectively, which is less negative than those values found for the cobalt(III) analogues. Trans-[Rh(en)2I2] · (ClO4) ammoniates totrans-[Rh(en)2(NH3)I] · (ClO4)2. The contribution of spontaneous ammoniation to the overall reaction oftrans-[Rh(en)2I2] · (ClO4) is negligible, so the uniqueness oftrans-[Co(en)2Cl2] · (ClO4) among cobalt(III) complexes in this respect is not reproduced for thetrans-dihalotetraamine structure in rhodium(III) complexes. A comparison of cobalt(III) and rhodium(III) amines with respect to activation parameters and the influence of formal charge of the metal complex on reactivity indicates a more associative type of activation for rhodium(III).  相似文献   

12.
Summary The oxidation of [Fe(phen)2(CN)2] and [Fe(bipy)2(CN)2] by nitrous acid in sulphuric acid follows the kinetic equation rate = k[H+] [HNO2] [complex] at low acidities. The mechanism is a diffusion controlled reaction between NO+ and the complex. Reaction is too slow for satisfactory use as a redox indicator for nitrite titrations at low acidities (0.1 M) [H+]. The variation of rate with acidity in more concentrated sulphuric acid (up to 6 M) is interpreted in terms of protonation of the complex to form [Fe(phen)2(CNH)2]2+.We thank the British Council for a maintenance award for P.R., and the Universidad Tecnica Federico Santa Maria, Valparaiso, Chile, for study leave.  相似文献   

13.
Density functional calculations for the [(RS)xFe(NO)4−x] (R=CH3) compounds are carried out using the DFT method with the B3LYP functional. The results can be verified by the experimental data only in the case of the [(RS)2Fe(NO)2] complex. The experimentally characterised molecular structure of [(RS)2Fe(NO)2] (where (RS)2=(SCH2CH2NMeCH2CH2CH2NMeCH2CH2S) is properly reproduced by the RB3LYP method. The discrepancy between the calculated spin densities with the integral spin observed experimentally is interpreted in terms of antiferromagnetic coupling between the Fe(III) centre and the NO ligands. The theoretical analysis gives a good account of some properties observed in these compounds. In particular, the electronic spectrum calculated by the TDDFT method for [(CH3S)2Fe(NO)2] is similar in shape to the experimental one, although is hypsochromically shifted. The LLCT (Sπ→π*NO), LMCT (Sπ→d) or (π*NO→d+Sπ→d) and MLCT (d→π*NO) transitions are mostly responsible for absorption of the [(RS)xFe(NO)4−x] complexes within UV-Vis. The chemical reactivity of [(RS)2Fe(NO)2] is interpreted basing on the calculated effect of a polar solvent on the ligand polarity and on the character of the HOMO and LUMO orbitals.  相似文献   

14.
Activation parameters have been obtained for the chelation of Mo(CO)5dpe (dpe = Ph2PCH2CH2PPh2) and of Mo(CO)5dmpe (dmpe = Me2PCH2CH2PMe2) to give cis-Mo(CO)4dpe and cis-Mo(CO)4dmpe respectively. The results are compared with those for the analogous chromium complexes and show that the enthalpy contribution determines the more rapid chelation in the molybdenum complexes. The preparation and properties of the chelate-bridged hetero-metallic complex (CO)5ModmpeMn(CO)4Br are reported. The reaction between Et4N[Mn(CO)4X2] (X = Cl, Br) and bidentate ligands dpe, dmpe and ape (ape = Ph2PCH2CH2AsPh2) in the presence of either silver(I) tetrafluoroborate or Et3OBF4 produces cis-Mn(CO)4X(bidentate) which is identified by infrared and mass spectrometry. At room temperature the Mn(CO)4X(bidentate) complex is rapidly converted to the chelated fac-Mn(CO)3X(bidentate) complex. The chelation process is approximately 104 times more rapid than in the isoelectronic chromium(O) complexes. The preparation and characterisation of fac-Mn(CO)3Br(dmpe), cis-Mn(CO)4Br(PMe3) and fac-Mn(CO)3Br(PMe3)2 are reported.  相似文献   

15.
The growth behavior of (GaAl) n (n = 1–12) and the chemisorptions of hydrogen on the ground state geometries have been studied with the three-parameter hybrid generalized gradient approximation due to Becke-Lee–Yang–Parr (B3LYP). The dissociation energy, the second-order energy differences, and the HOMO–LUMO gaps indicate that the magic numbers of the calculated (GaAl) n clusters are n = 4 and 6. To my knowledge, this is the first time that a systematic study of chemisorptions of hydrogen on gallium aluminum clusters. The onefold top site of aluminum atom is identified to be the most favorable chemisorptions site for one hydrogen chemisorptions on most (GaAl) n clusters. In general, dissociative chemisorptions of a hydrogen molecule on a top site of aluminum atom is found common for all sizes clusters considered here except for (GaAl) n (n = 1–3) clusters. The stability of the (GaAl) n H m complexes shows that both large second-order difference and large fragmentation energies for (GaAl)10H2 and (GaAl)11H2 make these species behaving like magic clusters.  相似文献   

16.
4‐Methylpyridinium cations, mpyH+, crystallized with complex aqua/chloro/lanthanoid(III) species for the gamut of the rare earth series, have ‘domains of existence’ defined for the following forms:
  • (a) The triclinic series (mpyH)2[{(H2O)3Cl3Ln(μ‐Cl)(2|2)}2], previously defined for the Ln = La–Nd members (those for La, Pr characterized by full structure determinations);
  • (b) The triclinic series (mpyH)2[(H2O)3LnCl4]Cl, previously defined for the Ln = Eu, Ho members by full structure determinations, is here augmented by the Ln = Nd, Sm examples (full structure determinations, the latter a new ‘light’ Ln extremum);
  • (c) A new monoclinic C2/c series (mpyH)2[(H2O)4LnCl3]Cl2, defined for the Ln = Er–Yb extrema (full structural characterizations for the Ln = Er, Tm, Yb members); the lanthanoid‐containing entity (which, with the cations, exhibits some disorder) is a neutral molecule of a new type. For the Ln = Lu case, a fully ordered derivative monoclinic C2 form has been obtained in a cell one half the size.
Other types have also been characterized, thus:
  • (i) For Ln = La, (mpyH)8[{(H2O)3Cl3La(μ‐Cl)(2|2)}2]‐[{(H2O)4Cl2La(μ‐Cl)2La(OH2)2Cl2(μ‐Cl)(2|2)}2]Cl4·6H2O·mpy has been defined by a full structural determination; the binuclear anion is similar to that in (a), albeit with some disorder, with the tetranuclear anion derivative of it.
  • (ii) For Ln = Lu, (mpyH)2[(H2O)5LuCl2]Cl3 is defined by a full structure determination.
  • (iii) For Ln = Y, (mpyH)2[(H2O)7YCl]Cl4 is defined by a full structure determination; here, and in (ii), the complex component is cationic.
  相似文献   

17.
Because of its unsaturated bonds, C60 is susceptible to polymerize into dimers. The implications of nitrogen doping on the geometrical and electronic structure of C60 dimers have been ambiguous for years. A quarter‐century after the discovery of azafullerene dimer (C59N)2, we reported its single crystallographic structure in 2019. Herein, the unambiguous crystal structure information of (C59N)2 is elucidated specifically, revealing that the inter‐cage C—C single bond length of (C59N)2 is comparable with that of an ordinary C(sp3)‐C(sp3) single bond, and that the most stable conformer of (C59N)2 is gauche‐conformer with a dihedral angle of 66°. To amend the structural deviations, geometrical structure of (C59N)2 is optimized by a B3LYP‐D3BJ function, which is proved to be more consistent with its single crystal structure than those by the commonly used B3LYP function. Moreover, the calculation method is also suitable for other representative fullerene dimers, such as (C60)2 and its divalent anion. Additionally, the dissociation of (C59N)2 at 473 K under mass spectrometric conditions suggests the inter‐cage C—C bond is relatively weaker than an ordinary C—C single bond, which can be explained by the interaction energies of inter‐cages.  相似文献   

18.
The structures of three newly synthesized phosphonate‐substituted polyoxotitanates are reported. The Ti/O core of [Ti4O(OEt)12(PhenylPO3)] ( 1 ) is the building block for two larger phosphonate‐substituted nanoclusters, [Ti25O26(OEt)36(PhenylPO3)6] ( 2 ) and [Ti26O26(OEt)39(PhenylPO3)6]Br ( 3 ). All compounds exhibit a not previously recognized triply bridging binding mode of the phosphonate anchor with short connecting Ti? O bonds, the average of which is 2.010(7) Å. Comparison with previously reported work suggests that the binding mode of the phosphonate anchor is strongly dependent on the structure of the underlying substrate.  相似文献   

19.
The electrochemical behavior of iron diimine complexes, (H3C?N=C(R)?C(R′)=N?CH3)3Fe(II) (R, R′=H,H;H, CH3; CH3, CH3), and (C5H4N?C(R1)=N(R2))3Fe(II) (R1, R2=H, CH3; CH3, CH3) on a platinum working electrode in acetonitrile is described, and compared to that of the parent aromatic complex, tris-(2,2′-bipyridine)Fe(II). One-electron reversible oxidations were found for all the compounds studied. The electrochemical reductions show 2–3 reduction waves in the potential range studied. Only for the complexes of mixed diimine ligands or 2,2′-bipyridine, a pre-adsorption wave is also observed. It is possible to stabilize low valence states with all ligands studied. A formal iron(I) state is described for the first time for all aliphatic diimine complexes, thus showing that the acceptor properties of the diimine complexes do not depend on the presence of the aromatic rings, but on the iron-diimine chromophore.  相似文献   

20.
The synthesis of three new polynuclear carbonyltrialkylphosphine complexes of Pd0 is described. The formulae Pd12(CO)15 (PEt3)7, Pd12 (CO)12 (PBu3)7 and Pd12(CO)17 (PBu3)5 are proposed for the new complexes.  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