10 research outputs found

    An Effective Route to Dinuclear Niobium and Tantalum Imido Complexes

    No full text
    Thermal treatment of the trichloro complexes [MCl<sub>3</sub>(NR)­py<sub>2</sub>] (R = <i>t</i>Bu, Xyl; M = Nb, Ta) (Xyl = 2,6-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub>) under vacuum affords the dinuclear imido species [MCl<sub>2</sub>(μ-Cl)­(NR)­py]<sub>2</sub> (R = <i>t</i>Bu, Xyl; M = Nb <b>1</b>, <b>3</b>; Ta <b>2</b>, <b>4</b>) with loss of pyridine. Complexes <b>1</b>–<b>4</b> can be easily transformed to the mononuclear starting materials [MCl<sub>3</sub>(NR)­py<sub>2</sub>] (R = <i>t</i>Bu, Xyl; M = Nb, Ta) upon reaction with pyridine. While reactions of compounds <b>1</b> and <b>2</b> with a series of alkylating reagents render the mononuclear peralkylated imido complexes [MR<sub>3</sub>(N<i>t</i>Bu)] (R = Me, CH<sub>2</sub>Ph, CH<sub>2</sub>CMe<sub>3</sub>, CH<sub>2</sub>CMePh, CH<sub>2</sub>SiMe<sub>3</sub>), the analogous treatment with allylmagnesium chloride results in the formation of the dinuclear niobium­(IV) derivative [(N<i>t</i>Bu)­(η<sup>3</sup>-C<sub>3</sub>H<sub>5</sub>)­M­(μ-C<sub>3</sub>H<sub>5</sub>)­(μ-Cl)<sub>2</sub>M­(N<i>t</i>Bu)­py<sub>2</sub>] (<b>5</b>). Additionally, the treatment of the starting materials <b>1</b> and <b>2</b> with the organosilicon reductant 1,4-bis­(trimethylsilyl)-1,4-diaza-2,5-cyclohexadiene yields the pyridyl-bridged dinuclear derivatives [M<sub>2</sub>Cl<sub>2</sub>(μ-Cl)<sub>2</sub>(N<i>t</i>Bu)<sub>2</sub>py<sub>2</sub>]<sub>2</sub>(μ-NC<sub>4</sub>H<sub>4</sub>N)<sub>2</sub> (M = Nb <b>6</b>, Ta <b>7</b>). Controlled hydrolysis reaction of <b>1</b> and <b>2</b> affords the oxo chlorido-bridged products [MCl­(μ-Cl)­(N<i>t</i>Bu)­py]<sub>2</sub>(μ-O) (M = Nb <b>8</b>, Ta <b>9</b>) in a quantitative way, while the treatment of these latter with one more equivalent of pyridine led to complexes [MCl<sub>2</sub>(N<i>t</i>Bu)­py<sub>2</sub>]<sub>2</sub>(μ-O) (M = Nb <b>10</b>, Ta <b>11</b>). Structural study of these dinuclear imido derivatives has been also performed by X-ray crystallography

    C–H Activation on an Oxo-Bridged Dititanium Complex: From Alkyl to μ‑Alkylidene Functionalities

    No full text
    Thermal treatment of the dinuclear compound [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(CH<sub>2</sub>SiMe<sub>3</sub>)<sub>2</sub>}<sub>2</sub>­(μ-O)] (<b>1</b>) provides the formation of the metallacycle derivatives [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>­(μ-CH<sub>2</sub>­SiMe<sub>2</sub>CH<sub>2</sub>)­(CH<sub>2</sub>­SiMe<sub>3</sub>)<sub>2</sub>­(μ-O)] (<b>2</b>) and [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>­(μ-CH<sub>2</sub>­SiMe<sub>2</sub>CH<sub>2</sub>)<sub>2</sub>­(μ-O)] (<b>3</b>) and the μ-alkylidene complex [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>­(μ-CH<sub>2</sub>­SiMe<sub>2</sub>CH<sub>2</sub>)­(μ-CHSiMe<sub>3</sub>)­(μ-O)] (<b>4</b>) by sequential carbon–hydrogen activation processes. The reaction of <b>3</b> with <i>tert</i>-butylisocyanide, in 1:1 and 1:2 ratios, leads to the η<sup>2</sup>-iminoacyl complexes [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>­(μ-<i>t</i>BuNC­CH<sub>2</sub>­SiMe<sub>2</sub>CH<sub>2</sub>)­(μ-CH<sub>2</sub>­SiMe<sub>2</sub>CH<sub>2</sub>)­(μ-O)] (<b>5</b>) and [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>­(μ-<i>t</i>BuNC­CH<sub>2</sub>­SiMe<sub>2</sub>CH<sub>2</sub>)<sub>2</sub>­(μ-O)] (<b>6</b>), respectively. The molecular structures of complexes <b>3</b>, <b>4</b>, <b>5</b>, and <b>6</b> have been determined by single-crystal X-ray diffraction analyses

    C–H Activation on an Oxo-Bridged Dititanium Complex: From Alkyl to μ‑Alkylidene Functionalities

    No full text
    Thermal treatment of the dinuclear compound [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(CH<sub>2</sub>SiMe<sub>3</sub>)<sub>2</sub>}<sub>2</sub>­(μ-O)] (<b>1</b>) provides the formation of the metallacycle derivatives [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>­(μ-CH<sub>2</sub>­SiMe<sub>2</sub>CH<sub>2</sub>)­(CH<sub>2</sub>­SiMe<sub>3</sub>)<sub>2</sub>­(μ-O)] (<b>2</b>) and [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>­(μ-CH<sub>2</sub>­SiMe<sub>2</sub>CH<sub>2</sub>)<sub>2</sub>­(μ-O)] (<b>3</b>) and the μ-alkylidene complex [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>­(μ-CH<sub>2</sub>­SiMe<sub>2</sub>CH<sub>2</sub>)­(μ-CHSiMe<sub>3</sub>)­(μ-O)] (<b>4</b>) by sequential carbon–hydrogen activation processes. The reaction of <b>3</b> with <i>tert</i>-butylisocyanide, in 1:1 and 1:2 ratios, leads to the η<sup>2</sup>-iminoacyl complexes [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>­(μ-<i>t</i>BuNC­CH<sub>2</sub>­SiMe<sub>2</sub>CH<sub>2</sub>)­(μ-CH<sub>2</sub>­SiMe<sub>2</sub>CH<sub>2</sub>)­(μ-O)] (<b>5</b>) and [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>­(μ-<i>t</i>BuNC­CH<sub>2</sub>­SiMe<sub>2</sub>CH<sub>2</sub>)<sub>2</sub>­(μ-O)] (<b>6</b>), respectively. The molecular structures of complexes <b>3</b>, <b>4</b>, <b>5</b>, and <b>6</b> have been determined by single-crystal X-ray diffraction analyses

    Systematic Approach for the Construction of Niobium and Tantalum Sulfide Clusters

    No full text
    Treatment of the imido complexes [MCl<sub>3</sub>(NR)­py<sub>2</sub>] (R = <sup><i>t</i></sup>Bu, 2,6-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub>; M = Nb <b>1</b>, <b>3</b>; Ta <b>2</b>, <b>4</b>) (Xyl = 2,6-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub>) with (Me<sub>3</sub>Si)<sub>2</sub>S in a 1:1 ratio afforded the new cube-type sulfide clusters [MCl­(NR)­py­(μ<sub>3</sub>-S)]<sub>4</sub> (R = <sup><i>t</i></sup>Bu, 2,6-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub>; M = Nb <b>5</b>, <b>7</b>; Ta <b>6</b>, <b>8</b>) with loss of Me<sub>3</sub>SiCl. Reactions of <b>5</b> and <b>6</b> with cyclopentadienyllithium in 1:4 ratio resulted in the rupture of the coordinative M–S bonds and the replacement of a pyridine molecule and a chlorine atom by an η<sup>5</sup>-cyclopentadienyl group in each metal center, affording the compounds [M­(η<sup>5</sup>-C<sub>5</sub>H<sub>5</sub>)­(N<sup><i>t</i></sup>Bu)­(μ-S)]<sub>4</sub> (M = Nb <b>9</b>, Ta <b>10</b>). These processes may develop through formation of the complexes [M<sub>4</sub>(η<sup>5</sup>-C<sub>5</sub>H<sub>5</sub>)<sub>2</sub>(μ-Cl)­(N<sup><i>t</i></sup>Bu)<sub>4</sub>py<sub>2</sub>(μ<sub>3</sub>-S)<sub>2</sub>­(μ-S)<sub>2</sub>]­(C<sub>5</sub>H<sub>5</sub>) (M = Nb <b>11</b>, Ta <b>12</b>), also obtained by reaction of <b>5</b> and <b>6</b> with cyclopentadienyllithium in 1:3 ratio. As further evidence, <b>11</b> and <b>12</b> led to complexes <b>9</b> and <b>10</b> by treatment with one more equivalent of the lithium reagent. The structural study of these metal sulfide clusters has been also performed by X-ray crystallography

    Carbon–Nitrogen Bond Construction and Carbon–Oxygen Double Bond Cleavage on a Molecular Titanium Oxonitride: A Combined Experimental and Computational Study

    No full text
    New carbon–nitrogen bonds were formed on addition of isocyanide and ketone reagents to the oxonitride species [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(μ-O)}<sub>3</sub>(μ<sub>3</sub>-N)] (<b>1</b>). Reaction of <b>1</b> with XylNC (Xyl = 2,6-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub>) in a 1:3 molar ratio at room temperature leads to compound [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(μ-O)}<sub>3</sub>­(μ-XylNCCNXyl)­(NCNXyl)] (<b>2</b>), after the addition of the nitrido group to one coordinated isocyanide and the carbon–carbon coupling of the other two isocyanide molecules have taken place. Thermolysis of <b>2</b> gives [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(μ-O)}<sub>3</sub>­(XylNCNXyl)­(CN)] (<b>3</b>) where the heterocumulene [XylNCCNXyl] moiety and the carbodiimido [NCNXyl] fragment in <b>2</b> have undergone net transformations. Similarly, <i>tert</i>-butyl isocyanide (<i>t</i>BuNC) reacts with the starting material <b>1</b> under mild conditions to give the paramagnetic derivative [{Ti<sub>3</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>3</sub>­(μ-O)<sub>3</sub>­(NCN<i>t</i>Bu)}<sub>2</sub>­(μ-CN)<sub>2</sub>] (<b>4</b>). However, compound <b>1</b> provides the oxo ketimide derivatives [{Ti<sub>3</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>3</sub>­(μ-O)<sub>4</sub>}­(NCRPh)] [R = Ph (<b>5</b>), <i>p</i>-Me­(C<sub>6</sub>H<sub>4</sub>) (<b>6</b>), <i>o</i>-Me­(C<sub>6</sub>H<sub>4</sub>) (<b>7</b>)] upon reaction with benzophenone, <i>p</i>-methylbenzophenone, and <i>o</i>-methylbenzophenone, respectively. In these reactions, the carbon–oxygen double bond is completely ruptured, leading to the formation of a carbon–nitrogen and two metal–oxygen bonds. The molecular structures of complexes <b>2</b>–<b>4</b>, <b>6</b>, and <b>7</b> were determined by single-crystal X-ray diffraction analyses. Density functional theory calculations were performed on the incorporation of isocyanides and ketones to the model complex [{Ti­(η<sup>5</sup>-C<sub>5</sub>H<sub>5</sub>)­(μ-O)}<sub>3</sub>­(μ<sub>3</sub>-N)] (<b>1H</b>). The mechanism involves the coordination of the substrates to one of the titanium metal centers, followed by an isomerization to place those substrates cis with respect to the apical nitrogen of <b>1H</b>, where carbon–nitrogen bond formation occurs with a low-energy barrier. In the case of aryl isocyanides, the resulting complex incorporates additional isocyanide molecules leading to a carbon–carbon coupling. With ketones, the high oxophilicity of titanium promotes the unusual total cleavage of the carbon–oxygen double bond

    Co-complexation of Lithium Gallates on the Titanium Molecular Oxide {[Ti(η<sup>5</sup>‑C<sub>5</sub>Me<sub>5</sub>)(μ-O)}<sub>3</sub>(μ<sub>3</sub>‑CH)]

    No full text
    Amide and lithium aryloxide gallates [Li<sup>+</sup>{RGaPh<sub>3</sub>}<sup>−</sup>] (R = NMe<sub>2</sub>, O-2,6-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub>) react with the μ<sub>3</sub>-alkylidyne oxoderivative ligand [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(μ-O)}<sub>3</sub>(μ<sub>3</sub>-CH)] (<b>1</b>) to afford the gallium–lithium–titanium cubane complexes [{Ph<sub>3</sub>Ga­(μ-R)­Li}­{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(μ-O)}<sub>3</sub>(μ<sub>3</sub>-CH)] [R = NMe<sub>2</sub> (<b>3</b>), O-2,6-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub> (<b>4</b>)]. The same complexes can be obtained by treatment of the [Ph<sub>3</sub>Ga­(μ<sub>3</sub>-O)<sub>3</sub>{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)}<sub>3</sub>(μ<sub>3</sub>-CH)] (<b>2</b>) adduct with the corresponding lithium amide or aryloxide, respectively. Complex <b>3</b> evolves with formation of <b>5</b> as a solvent-separated ion pair constituted by the lithium dicubane cationic species [Li­{(μ<sub>3</sub>-O)<sub>3</sub>Ti<sub>3</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>3</sub>(μ<sub>3</sub>-CH)}<sub>2</sub>]<sup>+</sup> together with the anionic [(GaPh<sub>3</sub>)<sub>2</sub>(μ-NMe<sub>2</sub>)]<sup>−</sup> unit. On the other hand, the reaction of <b>1</b> with Li­(<i>p</i>-MeC<sub>6</sub>H<sub>4</sub>) and GaPh<sub>3</sub> leads to the complex [Li­{(μ<sub>3</sub>-O)<sub>3</sub>Ti<sub>3</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>3</sub>(μ<sub>3</sub>-CH)}<sub>2</sub>]­[GaLi­(<i>p</i>-MeC<sub>6</sub>H<sub>4</sub>)<sub>2</sub>Ph<sub>3</sub>] (<b>6</b>). X-ray diffraction studies were performed on <b>1</b>, <b>2</b>, <b>4</b>, and <b>5</b>, while trials to obtain crystals of <b>6</b> led to characterization of [Li­{(μ<sub>3</sub>-O)<sub>3</sub>Ti<sub>3</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>3</sub>(μ<sub>3</sub>-CH)}<sub>2</sub>]­[PhLi­(μ-C<sub>6</sub>H<sub>5</sub>)<sub>2</sub>Ga­(<i>p</i>-MeC<sub>6</sub>H<sub>4</sub>)­Ph] <b>6a</b>

    Reactivity of Tuck-over Titanium Oxo Complexes with Isocyanides

    Get PDF
    The reactivity of the “tuck-over” species [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(CH<sub>2</sub>Ph)<sub>3</sub>­(μ-η<sup>5</sup>-C<sub>5</sub>Me<sub>4</sub>CH<sub>2</sub>-κ<i>C</i>)­(μ-O)] (<b>1</b>) and [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(CH<sub>2</sub>CMe<sub>3</sub>)­(μ-η<sup>5</sup>-C<sub>5</sub>Me<sub>4</sub>CH<sub>2</sub>-κ<i>C</i>)­(μ-CH<sub>2</sub>CMe<sub>2</sub>CH<sub>2</sub>)­(μ-O)] (<b>2</b>) toward isocyanides has been examined both synthetically and theoretically. Treatment of <b>1</b> with the isocyanides RNC, R = Me<sub>3</sub>SiCH<sub>2</sub>, 2,6-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub>, <i>t</i>Bu, <i>i</i>Pr, leads to a series of η<sup>2</sup>-iminoacyl species (<b>3</b>–<b>6</b>) where the molecule of isocyanide inserts into one of the terminal metal–alkyl bonds. The analogous reaction of the “tuck-over” metallacycle species <b>2</b> with 2,6-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub>NC and <i>t</i>BuNC results in the initial insertion of one isocyanide into the terminal Ti–alkyl bond to form the iminoacyl complexes <b>7</b> and <b>8</b>, followed by a second insertion into the metallacycle moiety to generate <b>9</b>, in the case of <i>tert</i>-butylisocyanide. DFT calculations support the selective reactivity observed experimentally with a kinetic and thermodynamic preference for RNC insertion on the terminal alkyl groups bound to both metallic centers over the alternative insertion on the “tuck-over” ligand

    Reactivity of Tuck-over Titanium Oxo Complexes with Isocyanides

    No full text
    The reactivity of the “tuck-over” species [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(CH<sub>2</sub>Ph)<sub>3</sub>­(μ-η<sup>5</sup>-C<sub>5</sub>Me<sub>4</sub>CH<sub>2</sub>-κ<i>C</i>)­(μ-O)] (<b>1</b>) and [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(CH<sub>2</sub>CMe<sub>3</sub>)­(μ-η<sup>5</sup>-C<sub>5</sub>Me<sub>4</sub>CH<sub>2</sub>-κ<i>C</i>)­(μ-CH<sub>2</sub>CMe<sub>2</sub>CH<sub>2</sub>)­(μ-O)] (<b>2</b>) toward isocyanides has been examined both synthetically and theoretically. Treatment of <b>1</b> with the isocyanides RNC, R = Me<sub>3</sub>SiCH<sub>2</sub>, 2,6-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub>, <i>t</i>Bu, <i>i</i>Pr, leads to a series of η<sup>2</sup>-iminoacyl species (<b>3</b>–<b>6</b>) where the molecule of isocyanide inserts into one of the terminal metal–alkyl bonds. The analogous reaction of the “tuck-over” metallacycle species <b>2</b> with 2,6-Me<sub>2</sub>C<sub>6</sub>H<sub>3</sub>NC and <i>t</i>BuNC results in the initial insertion of one isocyanide into the terminal Ti–alkyl bond to form the iminoacyl complexes <b>7</b> and <b>8</b>, followed by a second insertion into the metallacycle moiety to generate <b>9</b>, in the case of <i>tert</i>-butylisocyanide. DFT calculations support the selective reactivity observed experimentally with a kinetic and thermodynamic preference for RNC insertion on the terminal alkyl groups bound to both metallic centers over the alternative insertion on the “tuck-over” ligand

    Intermetallic Cooperation in C–H Activation Involving Transient Titanium-Alkylidene Species: A Synthetic and Mechanistic Study

    No full text
    Remote carbon–hydrogen activation on titanium dinuclear complexes [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­R<sub>2</sub>}<sub>2</sub>(μ-O)] [R = CH<sub>2</sub>SiMe<sub>3</sub> <b>2</b>, CH<sub>2</sub>CMe<sub>3</sub> <b>3</b>, and CH<sub>2</sub>Ph <b>5</b>) have been examined both synthetically and theoretically. While the thermal treatment of the oxoderivative [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(CH<sub>2</sub>SiMe<sub>3</sub>)<sub>2</sub>}<sub>2</sub>(μ-O)] (<b>2</b>) led to a series of metallacycle complexes (<b>2a</b>–<b>c</b>) by sequential carbon–hydrogen activation processes, [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(CH<sub>2</sub>CMe<sub>3</sub>)<sub>2</sub>}<sub>2</sub>(μ-O)] (<b>3</b>) gave rise to the formation of the metallacycle tuck-over species [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(μ-η<sup>5</sup>-C<sub>5</sub>Me<sub>4</sub>CH<sub>2</sub>-κC)­(CH<sub>2</sub>CMe<sub>3</sub>)­(μ-CH<sub>2</sub>CMe<sub>2</sub>CH<sub>2</sub>)­(μ-O)] (<b>4</b>), as result of hydrogen abstraction from a η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub> ligand. However, the thermolysis of the tetrabenzyl complex [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(CH<sub>2</sub>Ph)<sub>2</sub>}<sub>2</sub>(μ-O)] (<b>5</b>) yielded the derivative [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(μ-η<sup>5</sup>-C<sub>5</sub>Me<sub>4</sub>CH<sub>2</sub>-κC)­(CH<sub>2</sub>Ph)<sub>3</sub>(μ-O)] (<b>6</b>) that only exhibits tuck-over η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub> metalation. DFT calculations show that the mechanism involves a first α-hydrogen abstraction to generate a transient titanium alkylidene, which enables it to activate β- and γ-C­(sp<sup>3</sup>)-H bonds on the adjacent titanium center. The calculations also establish a reactivity order for the different type of γ-H abstractions, trimethylsilyl > neopentyl ≌ benzyl, allowing us to explain the observed selectivity

    Intermetallic Cooperation in C–H Activation Involving Transient Titanium-Alkylidene Species: A Synthetic and Mechanistic Study

    No full text
    Remote carbon–hydrogen activation on titanium dinuclear complexes [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­R<sub>2</sub>}<sub>2</sub>(μ-O)] [R = CH<sub>2</sub>SiMe<sub>3</sub> <b>2</b>, CH<sub>2</sub>CMe<sub>3</sub> <b>3</b>, and CH<sub>2</sub>Ph <b>5</b>) have been examined both synthetically and theoretically. While the thermal treatment of the oxoderivative [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(CH<sub>2</sub>SiMe<sub>3</sub>)<sub>2</sub>}<sub>2</sub>(μ-O)] (<b>2</b>) led to a series of metallacycle complexes (<b>2a</b>–<b>c</b>) by sequential carbon–hydrogen activation processes, [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(CH<sub>2</sub>CMe<sub>3</sub>)<sub>2</sub>}<sub>2</sub>(μ-O)] (<b>3</b>) gave rise to the formation of the metallacycle tuck-over species [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(μ-η<sup>5</sup>-C<sub>5</sub>Me<sub>4</sub>CH<sub>2</sub>-κC)­(CH<sub>2</sub>CMe<sub>3</sub>)­(μ-CH<sub>2</sub>CMe<sub>2</sub>CH<sub>2</sub>)­(μ-O)] (<b>4</b>), as result of hydrogen abstraction from a η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub> ligand. However, the thermolysis of the tetrabenzyl complex [{Ti­(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(CH<sub>2</sub>Ph)<sub>2</sub>}<sub>2</sub>(μ-O)] (<b>5</b>) yielded the derivative [Ti<sub>2</sub>(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)­(μ-η<sup>5</sup>-C<sub>5</sub>Me<sub>4</sub>CH<sub>2</sub>-κC)­(CH<sub>2</sub>Ph)<sub>3</sub>(μ-O)] (<b>6</b>) that only exhibits tuck-over η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub> metalation. DFT calculations show that the mechanism involves a first α-hydrogen abstraction to generate a transient titanium alkylidene, which enables it to activate β- and γ-C­(sp<sup>3</sup>)-H bonds on the adjacent titanium center. The calculations also establish a reactivity order for the different type of γ-H abstractions, trimethylsilyl > neopentyl ≌ benzyl, allowing us to explain the observed selectivity
    corecore