80 research outputs found

    <sup>29</sup>Si NMR Spectra of Silicon-Containing Uranium Complexes

    No full text
    <sup>29</sup>Si NMR spectra have been recorded for a series of uranium complexes containing silicon, and the data have been combined with results in the literature to determine if any trends exist between chemical shift and structure, ligand type, or oxidation state. Data on 52 paramagnetic inorganic and organometallic uranium complexes are presented. The survey reveals that, although there is some overlap in the range of shifts of U<sup>4+</sup> complexes versus U<sup>3+</sup> complexes, in general U<sup>3+</sup> species have shifts more negative than those of their U<sup>4+</sup> analogues. The single U<sup>2+</sup> example has the most negative shift of all at −322 ppm at 170 K. With only a few exceptions, U<sup>4+</sup> complexes have shifts between 0 and −150 ppm (vs SiMe<sub>4</sub>), whereas U<sup>3+</sup> complexes resonate between −120 and −250 ppm. The small data set on U<sup>5+</sup> species exhibits a broad 250 ppm range centered near 40 ppm. The data also show that aromatic ligands such as cyclopentadienide, cyclooctatetraenide, and the pentalene dianion exhibit chemical shifts less negative than those of other types of ligands

    Solvent-Free Organometallic Reactivity: Synthesis of Hydride and Carboxylate Complexes of Uranium and Yttrium from Gas/Solid Reactions

    No full text
    Gas/solid reactions involving H<sub>2</sub> and CO<sub>2</sub> with the metallocenes (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UMe<sub>2</sub> and (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­(allyl)<sub>2</sub> as solids in the absence of solvent provide an improved method to make organouranium hydride and carboxylate products. Decomposition products that can form in solution from the reactive hydrides can be avoided by this method, and this approach can also provide intermediates too reactive to isolate in some solution reactions. In contrast to the variable nature of the hydrogenolysis reaction of (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UMe<sub>2</sub> in toluene that forms byproducts along with the mixture of [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UH<sub>2</sub>]<sub>2</sub> and [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UH]<sub>2</sub>, a byproduct-free hydrogenolysis occurs when (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UMe<sub>2</sub> in the solid state is treated with H<sub>2</sub> gas to form predominantly [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UH<sub>2</sub>]<sub>2</sub>. H<sub>2</sub> reacts with solid (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­(C<sub>3</sub>H<sub>5</sub>)<sub>2</sub> and (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­(C<sub>3</sub>H<sub>5</sub>) similarly. The reaction of CO<sub>2</sub> (80 psi) with solid (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UMe<sub>2</sub> forms the monocarboxylate (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­(O<sub>2</sub>CCH<sub>3</sub>-κ<sup>2</sup><i>O</i>,<i>O</i>′)­Me, in contrast to the solution reaction that forms the diacetate (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­(O<sub>2</sub>CCH<sub>3</sub>-κ<sup>2</sup><i>O</i>,<i>O</i>′)<sub>2</sub> in minutes. The reaction of H<sub>2</sub> with solid (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>Y­(C<sub>3</sub>H<sub>5</sub>) provided (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>Y­(μ-H)­YH­(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub> without the decomposition products that it forms in solution such that single crystals suitable for X-ray diffraction could be isolated for the first time

    Solvent-Free Organometallic Reactivity: Synthesis of Hydride and Carboxylate Complexes of Uranium and Yttrium from Gas/Solid Reactions

    No full text
    Gas/solid reactions involving H<sub>2</sub> and CO<sub>2</sub> with the metallocenes (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UMe<sub>2</sub> and (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­(allyl)<sub>2</sub> as solids in the absence of solvent provide an improved method to make organouranium hydride and carboxylate products. Decomposition products that can form in solution from the reactive hydrides can be avoided by this method, and this approach can also provide intermediates too reactive to isolate in some solution reactions. In contrast to the variable nature of the hydrogenolysis reaction of (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UMe<sub>2</sub> in toluene that forms byproducts along with the mixture of [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UH<sub>2</sub>]<sub>2</sub> and [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UH]<sub>2</sub>, a byproduct-free hydrogenolysis occurs when (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UMe<sub>2</sub> in the solid state is treated with H<sub>2</sub> gas to form predominantly [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UH<sub>2</sub>]<sub>2</sub>. H<sub>2</sub> reacts with solid (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­(C<sub>3</sub>H<sub>5</sub>)<sub>2</sub> and (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­(C<sub>3</sub>H<sub>5</sub>) similarly. The reaction of CO<sub>2</sub> (80 psi) with solid (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UMe<sub>2</sub> forms the monocarboxylate (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­(O<sub>2</sub>CCH<sub>3</sub>-κ<sup>2</sup><i>O</i>,<i>O</i>′)­Me, in contrast to the solution reaction that forms the diacetate (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­(O<sub>2</sub>CCH<sub>3</sub>-κ<sup>2</sup><i>O</i>,<i>O</i>′)<sub>2</sub> in minutes. The reaction of H<sub>2</sub> with solid (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>Y­(C<sub>3</sub>H<sub>5</sub>) provided (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>Y­(μ-H)­YH­(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub> without the decomposition products that it forms in solution such that single crystals suitable for X-ray diffraction could be isolated for the first time

    Synthesis and CO<sub>2</sub> Insertion Reactivity of Allyluranium Metallocene Complexes

    No full text
    The U<sup>4+</sup> metallocene allyl chloride complexes (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[η<sup>3</sup>-CH<sub>2</sub>C­(R)­CH<sub>2</sub>]Cl (R = H, Me) can be synthesized by reaction of (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UCl<sub>2</sub> with 1 equiv of the corresponding allyl Grignard reagents, [CH<sub>2</sub>C­(R)­CH<sub>2</sub>]­MgCl, in hydrocarbon solvents. Bis­(allyl)­uranium complexes can also be obtained in this manner using 2 equiv of the corresponding allyl Grignard, and X-ray crystallographic studies reveal the presence of both η<sup>3</sup>- and η<sup>1</sup>-allyl ligands: (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[η<sup>3</sup>-CH<sub>2</sub>C­(R)­CH<sub>2</sub>]­[η<sup>1</sup>-CH<sub>2</sub>C­(R)CH<sub>2</sub>]. Sodium amalgam reduction of (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[η<sup>3</sup>-CH<sub>2</sub>C­(R)­CH<sub>2</sub>]Cl generates the U<sup>3+</sup> metallocene allyl complexes (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[η<sup>3</sup>-CH<sub>2</sub>C­(R)­CH<sub>2</sub>]. Carbon dioxide reacts with the U<sup>4+</sup> allyl complexes to form the U–C insertion products (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[κ<sup>2</sup><i>O</i>,<i>O</i>′-O<sub>2</sub>CCH<sub>2</sub>CHCH<sub>2</sub>]<sub>2–<i>x</i></sub>Cl<sub>x</sub> (<i>x</i> = 0, 1). The dicarboxylate (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[κ<sup>2</sup><i>O</i>,<i>O</i>′-O<sub>2</sub>CCH<sub>2</sub>CHCH<sub>2</sub>]<sub>2</sub>, which has a 171.98(5)° O–U–O angle, reacts with Me<sub>3</sub>SiCl to regenerate (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>UCl<sub>2</sub> and liberate Me<sub>3</sub>SiOC­(O)­CH<sub>2</sub>CHCH<sub>2</sub>

    Reactivity of U<sup>3+</sup> Metallocene Allyl Complexes Leads to a Nanometer-Sized Uranium Carbonate, [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U]<sub>6</sub>(μ‑κ<sup>1</sup>:κ<sup>2</sup>‑CO<sub>3</sub>)<sub>6</sub>

    No full text
    The U<sup>3+</sup> allyl complexes (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[CH<sub>2</sub>C­(R)­CH<sub>2</sub>] (R = H, Me) display three different types of reactivity, as exemplified by reactions with PhNNPh, cyclooctatetraene, and CO<sub>2</sub>. Two equivalents of (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[CH<sub>2</sub>C­(R)­CH<sub>2</sub>] effect a four-electron reduction of PhNNPh to form the bis­(imido) complex (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­(NPh)<sub>2</sub> and the bis­(allyl) species (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[CH<sub>2</sub>C­(R)­CH<sub>2</sub>]<sub>2</sub>. Two-electron reduction of C<sub>8</sub>H<sub>8</sub> occurs to form (C<sub>5</sub>Me<sub>5</sub>)­(C<sub>8</sub>H<sub>8</sub>)­U­[CH<sub>2</sub>C­(R)­CH<sub>2</sub>] products that contain only one cyclopentadienyl ring per metal. With CO<sub>2</sub> at 80 psi, both reduction and insertion occur. A hexametallic uranium carbonate, [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U]<sub>6</sub>(μ-κ<sup>1</sup>:κ<sup>2</sup>-CO<sub>3</sub>)<sub>6</sub>, is isolated as well as the bis­(carboxylate) complexes (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[κ<sup>2</sup>-<i>O</i>,<i>O</i>′-O<sub>2</sub>CCH<sub>2</sub>C­(R)CH<sub>2</sub>]<sub>2</sub>. The polymetallic carbonate complex can also be synthesized from [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U]<sub>2</sub>(μ-η<sup>6</sup>:η<sup>6</sup>-C<sub>6</sub>H<sub>6</sub>) and [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U]­[(μ-Ph)<sub>2</sub>BPh<sub>2</sub>] and CO<sub>2</sub>

    Reactivity of U<sup>3+</sup> Metallocene Allyl Complexes Leads to a Nanometer-Sized Uranium Carbonate, [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U]<sub>6</sub>(μ‑κ<sup>1</sup>:κ<sup>2</sup>‑CO<sub>3</sub>)<sub>6</sub>

    No full text
    The U<sup>3+</sup> allyl complexes (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[CH<sub>2</sub>C­(R)­CH<sub>2</sub>] (R = H, Me) display three different types of reactivity, as exemplified by reactions with PhNNPh, cyclooctatetraene, and CO<sub>2</sub>. Two equivalents of (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[CH<sub>2</sub>C­(R)­CH<sub>2</sub>] effect a four-electron reduction of PhNNPh to form the bis­(imido) complex (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­(NPh)<sub>2</sub> and the bis­(allyl) species (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[CH<sub>2</sub>C­(R)­CH<sub>2</sub>]<sub>2</sub>. Two-electron reduction of C<sub>8</sub>H<sub>8</sub> occurs to form (C<sub>5</sub>Me<sub>5</sub>)­(C<sub>8</sub>H<sub>8</sub>)­U­[CH<sub>2</sub>C­(R)­CH<sub>2</sub>] products that contain only one cyclopentadienyl ring per metal. With CO<sub>2</sub> at 80 psi, both reduction and insertion occur. A hexametallic uranium carbonate, [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U]<sub>6</sub>(μ-κ<sup>1</sup>:κ<sup>2</sup>-CO<sub>3</sub>)<sub>6</sub>, is isolated as well as the bis­(carboxylate) complexes (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U­[κ<sup>2</sup>-<i>O</i>,<i>O</i>′-O<sub>2</sub>CCH<sub>2</sub>C­(R)CH<sub>2</sub>]<sub>2</sub>. The polymetallic carbonate complex can also be synthesized from [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U]<sub>2</sub>(μ-η<sup>6</sup>:η<sup>6</sup>-C<sub>6</sub>H<sub>6</sub>) and [(C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>U]­[(μ-Ph)<sub>2</sub>BPh<sub>2</sub>] and CO<sub>2</sub>

    Expanding Yttrium Bis(trimethylsilylamide) Chemistry Through the Reaction Chemistry of (N<sub>2</sub>)<sup>2–</sup>, (N<sub>2</sub>)<sup>3–</sup>, and (NO)<sup>2–</sup> Complexes

    No full text
    The reaction chemistry of the side-on bound (N<sub>2</sub>)<sup>2–</sup>, (N<sub>2</sub>)<sup>3–</sup>, and (NO)<sup>2–</sup> complexes of the [(R<sub>2</sub>N)<sub>2</sub>Y]<sup>+</sup> cation (R = SiMe<sub>3</sub>), namely, [(R<sub>2</sub>N)<sub>2</sub>(THF)­Y]<sub>2</sub>(<i>μ</i>-η<sup>2</sup>:η<sup>2</sup>-N<sub>2</sub>), <b>1</b>, [(R<sub>2</sub>N)<sub>2</sub>(THF)­Y]<sub>2</sub>(<i>μ</i>-η<sup>2</sup>:η<sup>2</sup>-N<sub>2</sub>)­K, <b>2</b>, and [(R<sub>2</sub>N)<sub>2</sub>(THF)­Y]<sub>2</sub>(<i>μ</i>-η<sup>2</sup>:η<sup>2</sup>-NO), <b>3</b>, with oxidizing agents has been explored to search for other (E<sub>2</sub>)<sup><i>n</i>−</sup>, (E = N, O), species that can be stabilized by this cation. This has led to the first examples for the [(R<sub>2</sub>N)<sub>2</sub>Y]<sup>+</sup> cation of two fundamental classes of [(monoanion)<sub>2</sub>Ln]<sup>+</sup> rare earth systems (Ln = Sc, Y, lanthanides), namely, oxide complexes and the tetraphenylborate salt. In addition, an unusually high yield reaction with dioxygen was found to give a peroxide complex that completes the (N<sub>2</sub>)<sup>2–</sup>, (NO)<sup>2–</sup>, (O<sub>2</sub>)<sup>2–</sup> series with <b>1</b> and <b>3</b>. Specifically, the (<i><i>μ</i>-</i>O)<sup>2–</sup> oxide-bridged bimetallic complex, [(R<sub>2</sub>N)<sub>2</sub>(THF)­Y}<sub>2</sub>(<i><i>μ</i>-</i>O), <b>4</b>, is obtained as a byproduct from reactions of either the (N<sub>2</sub>)<sup>2–</sup> complex, <b>1</b>, or the (N<sub>2</sub>)<sup>3–</sup> complex, <b>2</b>, with NO, while the oxide formed from <b>2</b> with N<sub>2</sub>O is a polymeric species incorporating potassium, {[(R<sub>2</sub>N)<sub>2</sub>Y]<sub>2</sub>(<i><i>μ</i>-</i>O)<sub>2</sub>K<sub>2</sub>(<i><i>μ</i>-</i>C<sub>7</sub>H<sub>8</sub>)}<sub><i>n</i></sub>, <b>5</b>. Reaction of <b>1</b> with 1 atm of O<sub>2</sub> generates the (O<sub>2</sub>)<sup>2–</sup> bridging side-on peroxide [(R<sub>2</sub>N)<sub>2</sub>(THF)­Y]<sub>2</sub>(<i>μ</i>-η<sup>2</sup>:η<sup>2</sup>-O<sub>2</sub>), <b>6</b>. The O–O bond in <b>6</b> is cleaved by KC<sub>8</sub> to provide an alternative synthetic route to <b>5</b>. Attempts to oxidize the (NO)<sup>2–</sup> complex, <b>3</b>, with AgBPh<sub>4</sub> led to the isolation of the tetraphenylborate complex, [(R<sub>2</sub>N)<sub>2</sub>Y­(THF)<sub>3</sub>]­[BPh<sub>4</sub>], <b>7</b>, that was also synthesized from <b>1</b> and AgBPh<sub>4</sub>. Oxidation of the (N<sub>2</sub>)<sup>2–</sup> complex, <b>1</b>, with the radical trap (2,2,6,6-tetramethylpiperidin-1-yl)­oxyl, TEMPO, generates the (TEMPO)<sup>−</sup> anion complex, (R<sub>2</sub>N)<sub>2</sub>(THF)­Y­(η<sup>2</sup>-ONC<sub>5</sub>H<sub>6</sub>Me<sub>4</sub>), <b>8</b>

    Synthesis and Structure of Bis- and Tris-Benzyl Bismuth Complexes

    No full text
    The first crystallographic characterization of bismuth complexes containing benzyl ligands is reported. The NCN pincer ligand complex, Ar′BiCl<sub>2</sub> [Ar′ = 2,6-(Me<sub>2</sub>NCH<sub>2</sub>)<sub>2</sub>C<sub>6</sub>H<sub>3</sub>], reacts with (PhCH<sub>2</sub>)­MgCl to form Ar′Bi­(η<sup>1</sup>-CH<sub>2</sub>Ph)<sub>2</sub> in high yield. X-ray crystallography and spectroscopic studies confirm η<sup>1</sup>-bonding of the benzyl ligands in Ar′Bi­(η<sup>1</sup>-CH<sub>2</sub>Ph)<sub>2</sub> as well as in the homoleptic Bi­(η<sup>1</sup>-CH<sub>2</sub>Ph)<sub>3</sub>

    Cocrystallization of (μ‑S<sub>2</sub>)<sup>2–</sup> and (μ-S)<sup>2–</sup> and Formation of an [η<sup>2</sup>‑S<sub>3</sub>N(SiMe<sub>3</sub>)<sub>2</sub>] Ligand from Chalcogen Reduction by (N<sub>2</sub>)<sup>2–</sup> in a Bimetallic Yttrium Amide Complex

    No full text
    The reactivity of the (N<sub>2</sub>)<sup>2–</sup> complex {[(Me<sub>3</sub>Si)<sub>2</sub>N]<sub>2</sub>Y­(THF)}<sub>2</sub>(μ-η<sup>2</sup>:η<sup>2</sup>-N<sub>2</sub>) (<b>1</b>) with sulfur and selenium has been studied to explore the special reductive chemistry of this complex and to expand the variety of bimetallic rare-earth amide complexes. Complex <b>1</b> reacts with elemental sulfur to form a mixture of compounds, <b>2</b>, that is the first example of cocrystallized complexes of (S<sub>2</sub>)<sup>2–</sup> and S<sup>2–</sup> ligands. The crystals of <b>2</b> contain both the (μ-S<sub>2</sub>)<sup>2–</sup> complex {[(Me<sub>3</sub>Si)<sub>2</sub>N]<sub>2</sub>Y­(THF)}<sub>2</sub>(μ-η<sup>2</sup>:η<sup>2</sup>-S<sub>2</sub>) (<b>3</b>) and the (μ-S)<sup>2–</sup> complex {[(Me<sub>3</sub>Si)<sub>2</sub>N]<sub>2</sub>Y­(THF)}<sub>2</sub>(μ-S) (<b>4</b>), respectively. Modeling of the crystal data of <b>2</b> shows a 9:1 ratio of <b>3</b>:<b>4</b> in the crystals of <b>2</b> obtained from solutions that have 1:1 to 4:1 ratios of <b>3</b>/<b>4</b> by <sup>1</sup>H NMR spectroscopy. The addition of KC<sub>8</sub> to samples of <b>2</b> allows for the isolation of single crystals of <b>4</b>. The [S<sub>3</sub>N­(SiMe<sub>3</sub>)<sub>2</sub>]<sup>−</sup> ligand was isolated for the first time in crystals of [(Me<sub>3</sub>Si)<sub>2</sub>N]<sub>2</sub>Y­[η<sup>2</sup>-S<sub>3</sub>N­(SiMe<sub>3</sub>)<sub>2</sub>]­(THF) (<b>5</b>), obtained from the mother liquor of <b>2</b>. In contrast to the sulfur chemistry, the (μ-Se<sub>2</sub>)<sup>2–</sup> analogue of <b>3</b>, namely, {[(Me<sub>3</sub>Si)<sub>2</sub>N]<sub>2</sub>Y­(THF)}<sub>2</sub>(μ-η<sup>2</sup>:η<sup>2</sup>-Se<sub>2</sub>) (<b>6</b>), can be cleanly synthesized in good yield by reacting <b>1</b>, with elemental selenium. The (μ-Se)<sup>2–</sup> analogue of <b>4</b>, namely, {[(Me<sub>3</sub>Si)<sub>2</sub>N]<sub>2</sub>Y­(THF)}<sub>2</sub>(μ-Se) (<b>7</b>), was synthesized from Ph<sub>3</sub>PSe

    Reactivity of Organothorium Complexes with TEMPO

    No full text
    Reactions of the 2,2,6,6-tetramethylpiperidin-1-oxyl radical (TEMPO) with thorium metallocenes have been examined to investigate both the radical reaction patterns for organothorium complexes and the coordination chemistry of TEMPO with thorium. (η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>ThMe<sub>2</sub> reacts with 2 equiv of TEMPO to generate 1-methoxy-2,2,6,6-tetramethylpiperidine (Me-TEMPO) and (η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>ThMe­(η<sup>1</sup>-TEMPO), which contains a TEMPO<sup>–</sup> anion coordinated to thorium through oxygen only. (η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>Th­(η<sup>1</sup>-C<sub>3</sub>H<sub>5</sub>)­(η<sup>3</sup>-C<sub>3</sub>H<sub>5</sub>), synthesized from (η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>ThBr<sub>2</sub> and (C<sub>3</sub>H<sub>5</sub>)­MgBr, reacts with 2 equiv of TEMPO to form 1-(2-propen-1-yloxy)-2,2,6,6-tetramethylpiperidine (allyl-TEMPO) and (η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>Th­(η<sup>1</sup>-C<sub>3</sub>H<sub>5</sub>)­(η<sup>1</sup>-TEMPO). Although bis­(TEMPO) metallocenes were not obtained in these reactions, the methyl group in (η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>ThMe­(η<sup>1</sup>-TEMPO) is reactive with 1 equiv of CuBr to form (η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>ThBr­(η<sup>1</sup>-TEMPO). The bis­(TEMPO) metallocene (η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>Th­(η<sup>1</sup>-TEMPO)<sub>2</sub> is accessible in the reaction of [(η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>ThH<sub>2</sub>]<sub>2</sub> with 4 equiv of TEMPO. In contrast, (η<sup>5</sup>-C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub>ThBr<sub>2</sub> reacts with 2 equiv of TEMPO by loss of C<sub>5</sub>Me<sub>5</sub> to form (C<sub>5</sub>Me<sub>5</sub>)<sub>2</sub> and (η<sup>2</sup>-TEMPO)<sub>2</sub>ThBr<sub>2</sub>, in which the TEMPO<sup>–</sup> anions bind through oxygen and nitrogen. The bromide ions in (η<sup>2</sup>-TEMPO)<sub>2</sub>ThBr<sub>2</sub> can be replaced by an additional 2 equiv of TEMPO in the presence of 2 equiv of KC<sub>8</sub> to form the per­(TEMPO) complex Th­(η<sup>1</sup>-TEMPO)<sub>2</sub>(η<sup>2</sup>-TEMPO)<sub>2</sub>. ThBr<sub>4</sub>(THF)<sub>4</sub> reacts with TEMPO to form ThBr<sub>4</sub>(THF)<sub>2</sub>(η<sup>1</sup>-TEMPO), which contains an oxygen-bound TEMPO radical. The Th<sup>3+</sup> complex (η<sup>5</sup>-C<sub>5</sub>Me<sub>4</sub>H)<sub>3</sub>Th is oxidized in the presence of TEMPO, without ligand loss, to afford the Th<sup>4+</sup> species (η<sup>5</sup>-C<sub>5</sub>Me<sub>4</sub>H)<sub>3</sub>Th­(η<sup>1</sup>-TEMPO). The reactions show that TEMPO can react with organothorium complexes in several ways including coordination, anion substitution, and cyclopentadienyl replacement
    • …
    corecore