Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 8 de 8
Filter
Add more filters










Database
Language
Publication year range
1.
J Am Chem Soc ; 131(17): 6154-60, 2009 May 06.
Article in English | MEDLINE | ID: mdl-19366267

ABSTRACT

The experimental electron density study of Ti(C(5)H(4)Me)(2)[(CH(2))(2)CMe(2)] provides direct evidence for the presence of (C-C)-->Ti agostic interactions. In accord with the model of Scherer and McGrady, the C(alpha)-C(beta) bond densities no longer show cylindrical symmetry in the vicinity of the Ti atom and differ markedly from those of the other C-C bonds. At the points along the C(alpha)-C(beta) bond where the deviation is maximal the electron density is elongated toward the metal center. The distortion is supported by parallel theoretical calculations. A calculation on an Mo complex in which the agostic interaction is absent supports the Scherer and McGrady criterion for agostic interactions. Despite the formal d(0) electron configuration for this Ti(IV) species, a significant nonzero population is observed for the d orbitals, the d orbital population is largest for the d(xy) orbital, the lobes of which point toward the two C(alpha) atoms. Of the three different basis sets for the Ti atom used in theoretical calculations with the B3LYP functional, only the 6-311++G** set for Ti agrees well with the experimental charge density distribution in the Ti-(C(alpha)-C(beta))(2) plane.

2.
Magn Reson Chem ; 45(5): 393-400, 2007 May.
Article in English | MEDLINE | ID: mdl-17394221

ABSTRACT

This paper presents novel measurements and calculations of the olefinic (13)C chemical shift tensor principal values in several metal diene complexes. The experimental values and the calculations show shifts as large as 70 ppm with respect to the values in the parent olefinic compounds. These shifts are highly anisotropic, with the largest ones observed in the less shielded principal components and the smallest ones in the most shielded principal components of the tensor. The orientations of the principal components of the tensors remain, within 10 degrees , at their directions in ethylene and other olefinic compounds. The calculations, performed using the GIAO method and the LanDZ pseudopotential basis set, show good agreement with the experiments, and were used to establish definite evidence for the existence of a Cl-bridge structure in the bicyclo[2.2.1]hepta-2,5-diene (BCHD)dichlororuthenium(II) polymer.


Subject(s)
Alkadienes/chemistry , Magnetic Resonance Spectroscopy/methods , Metals, Heavy/chemistry , Norbornanes/chemistry , Organometallic Compounds/chemistry , Anisotropy , Carbon Isotopes , Ligands , Molybdenum/chemistry , Palladium/chemistry , Platinum/chemistry , Quantum Theory , Ruthenium/chemistry
3.
J Am Chem Soc ; 127(47): 16426-35, 2005 Nov 30.
Article in English | MEDLINE | ID: mdl-16305228

ABSTRACT

The reaction of the half-open titanocene, Ti(C5H5)(c-C8H11)(PMe3) (c-C8H11 = cyclooctadienyl), with two equivalents of PhC2SiMe3 leads to their incorporation and coupling to the dienyl fragment. One alkyne inserts into a C-H bond of the central CH2 group of the c-C8H11 ligand's edge-bridge, while the second undergoes a 5+2 coupling with the dienyl fragment, yielding coordinated sigma-allyl and olefin fragments, as demonstrated by X-ray diffraction. Together with the C5H5 and PMe3 coordinations, this leads to a 14-electron count. While the very electron-deficient titanium center passes up potential pi coordination of the allyl fragment, it instead engages in interactions with one or two C-C bonds, and perhaps a C-H bond, as revealed from the structural and spectroscopic data. Similar interactions have been found in electron-deficient metallacyclobutane complexes of titanium and zirconium, but not in the 18-electron molybdenum and tungsten analogues. These and other observations may have implications relating to metatheses and polymerizations of olefins.

4.
J Am Chem Soc ; 127(34): 11924-5, 2005 Aug 31.
Article in English | MEDLINE | ID: mdl-16117515

ABSTRACT

The normally remarkably inert SF6 has been found to be quite reactive toward low valent organometallic compounds, under conditions in which usually powerful fluorinating agents may be less reactive. Reaction of SF6 with Ti[1,3-C5H3(t-Bu)2](6,6-dmch)(PMe3), for example, leads to {Ti[1,3-C5H3(t-Bu)2]F2}4 (dmch = dimethylcyclohexadienyl), whose structure is based on a cube of fluoride ions with the ligated titanium centers situated above four coplanar face centers.

5.
Chem Commun (Camb) ; (11): 1462-4, 2005 Mar 21.
Article in English | MEDLINE | ID: mdl-15756336

ABSTRACT

The silylation of SBA-15 enhances the reducibility of cobalt oxides on a SBA-15 supported cobalt catalyst, and consequently increases the catalytic activity for Fischer-Tropsch synthesis of hydrocarbons from syngas and selectivity for longer chain products.

6.
J Am Chem Soc ; 126(43): 14105-16, 2004 Nov 03.
Article in English | MEDLINE | ID: mdl-15506775

ABSTRACT

Molecules of the form Cp(6,6-dmch)ZrX(2) (Cp = eta(5)-cyclopentadienyl, X = Cl, Br, I; 6,6-dmch = eta(5)-6,6-dimethylcyclohexadienyl) have been synthesized, and the molecular and electronic structures have been investigated. These molecules allow direct comparison of the bonding and properties of pentadienyl and cyclopentadienyl ligands in the same high-oxidation-state metal complexes. Unlike the well-known Cp(2)ZrX(2) analogues, these Cp(6,6-dmch)ZrX(2) molecules are intensely colored, indicating significantly different relative energies of the frontier orbitals. Also unusual, the average Zr-C distances to the 6,6-dmch pentadienyl ligand are about 0.1 A longer than the average Zr-C distances to the cyclopentadienyl ligand for these Zr(IV) complexes, opposite of what is observed for the Zr(II) complex Cp(2,6,6-tmch)Zr(PMe(3))(2) (tmch = eta(5)-2,6,6-trimethylcyclohexadienyl), reflecting a dramatic reversal in the favorability of the bonding depending on the metal oxidation state. The experimental and computational results indicate that the color of the Cp(6,6-dmch)ZrX(2) complexes is due to a 6,6-dmch ligand-to-metal charge-transfer band. Compared to the Cp(2)ZrX(2) analogues, the Cp(6,6-dmch)ZrX(2) molecules have a considerably less stable HOMO that is pentadienyl-based and an essentially unchanged metal-based LUMO. Also, the lowest unoccupied orbital of pentadienyl is stabilized relative to cyclopentadienyl and becomes a better potential delta electron acceptor, thus contributing to the differences in structure and reactivity of the low-valent and high-valent metal complexes.

7.
Dalton Trans ; (8): 1221-6, 2004 Apr 21.
Article in English | MEDLINE | ID: mdl-15252664

ABSTRACT

The synthesis and characterization of the open hafnocene, Hf(2,4-C7H11)2(PMe3)(C7H11 = dimethylpentadienyl), is reported. Additionally, a much improved synthetic procedure has been developed for Hf(2,4-C7H11)2(PEt3). Structural data have been obtained for these complexes, and for Ti(2,4-C7H11)2(PEt3) and Zr(2,4-C7H11)2(PMe3), thereby allowing for detailed comparisons between all M(2,4-C7H11)2(PX3) species (M = Ti, Zr, Hf; X = Me, Et). The presence of the coordinated phosphines led in all cases to the adoption of the expected syn-eclipsed geometries, with the phosphines positioned by the open dienyl edges. These phosphine ligands lead to substantial alterations of the bonding patterns in these species, relative to ligand-free complexes. Most notably, the shortest M-C distances involve the central dienyl carbon atoms. Additionally, the data reveal high degrees of steric crowding within these complexes, especially for the weakly bound Ti(2,4-C7H11)2(PEt3), and also demonstrate that significant deformations which have taken place within the dienyl ligands were substantially determined by the relative sizes of the metal centers.

8.
J Am Chem Soc ; 125(7): 1937-49, 2003 Feb 19.
Article in English | MEDLINE | ID: mdl-12580621

ABSTRACT

The crystal structure and electron density (ED) distribution of an imine coupling product with an open zirconocene, Zr(2,4-C(7)H(11))[(i-Pr)NCHPhCH(2)CMe=CHCMe=CH(2)] (C(7)H(11) = dimethylpentadienyl), have been derived from accurate synchrotron X-ray diffraction measurements. The molecular structure reveals asymmetric coordination of Zr by the pentadienyl (2,4-C(7)H(11)) ligand ( = 2.56(6) A), the butadiene fragment ( = 2.43(5) A), and the amide nitrogen atom (Zr-N = 2.0312(5) A) of the second ligand. The study of the ED and its topological analysis affords new insight into the bonding and electronic structure of the title zirconium complex. The interactions between the metal center and the ligands are represented by a Zr-N bond path and one Zr-C bond path with each of the pentadienyl and butadiene moieties, contrary to the usually depicted global metal-ligand bonding. The butadiene and pentadienyl groups exhibit a polarization of the corresponding pi-like ED in the C-Zr directions, indicating that the whole conjugated systems are nonetheless involved. The 4d atomic orbitals of Zr exhibit unusual populations according to ligand field considerations, which reveal a high degree of sigma-donation from the conjugated pi systems of both ligands. As deduced from numerical integration over the topologically defined atomic basins, the Zr to ligand charge transfer is 1.48 e to the C(17)NH(24) ligand and 0.68 e to dimethylpentadienyl. Topological analysis of a short intramolecular Zr.(C,H) contact provides no indication of the presence of agostic interactions, despite a small Zr-N-C angle of 102.87(4) degrees. Thus, no bond path and BCP (bond critical point) of the ED are found in the Zr.(C,H) region, which would have provided evidence for such direct interactions, nor is there any evidence for charge accumulation between the Zr and H atoms, or for lengthening of the C-H bond involved in the putative interaction. These characteristics, similar to those in other distorted situations, may be common for other electron-deficient d(0) complexes.

SELECTION OF CITATIONS
SEARCH DETAIL
...