Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 12 de 12
Filtrar
Más filtros











Base de datos
Intervalo de año de publicación
1.
Inorg Chem ; 53(22): 12066-75, 2014 Nov 17.
Artículo en Inglés | MEDLINE | ID: mdl-25347523

RESUMEN

The synthesis and characterization of Ln(Tp(iPr2))2 (Ln = Sm, 3Sm; Tm, 3Tm) are reported. While the simple (1)H NMR spectra of the compounds indicate a symmetrical solution structure, with equivalent pyrazolyl groups, the solid-state structure revealed an unexpected, "bent sandwich-like" geometry. By contrast, the structure of the less sterically congested Tm(Tp(Me2,4Et))2 (4) adopts the expected symmetrical structure with a linear B-Tm-B arrangement. Computational studies to investigate the origin of the unexpected bent structure of the former compounds indicate that steric repulsion between the isopropyl groups forces the Tp ligands apart and permits the development of unusual interligand C-H···N hydrogen-bonding interactions that help stabilize the structure. These results find support in the similar geometry of the Tm(III) analogue [Tm(Tp(iPr2))2]I, 3Tm(+), and confirm that the low symmetry is not the result of a metal-ligand interaction. The relevance of these results to the general question of the coordination geometry of MX2 and M(C5R5)2 (M = heavy alkaline earth and Ln(II), X = halide, and C5R5 = bulky persubstituted cyclopentadienyl) complexes and the importance of secondary H-bonding and nonbonding interactions on the structure are highlighted.

2.
Chemistry ; 18(25): 7886-95, 2012 Jun 18.
Artículo en Inglés | MEDLINE | ID: mdl-22573516

RESUMEN

Reaction mechanisms for the oxidative reactions of CO(2) and COS with [(C(5)Me(5))(2)Sm] have been investigated by means of DFT methods. The experimental formation of oxalate and dithiocarbonate complexes is explained. Their formation involve the samarium(III) bimetallic complexes [(C(5)Me(5))(2)Sm-CO(2)-Sm(C(5)Me(5))(2)] and [(C(5)Me(5))(2)Sm-COS-Sm(C(5)Me(5))(2)] as intermediates, respectively, ruling out radical coupling for the formation of the oxalate complex.

3.
Chem Commun (Camb) ; 47(44): 12203-5, 2011 Nov 28.
Artículo en Inglés | MEDLINE | ID: mdl-21987051

RESUMEN

The new divalent thulium compound [Tm(BH(4))(2)(DME)(2)] could be prepared by reduction of [Tm(BH(4))(3)(THF)(3)] or from TmI(2) and KBH(4). It was used as a precursor to the divalent [(Tp(tBu,Me))Tm(BH(4))(THF)] by reaction with potassium tris(2-tBu-4-Me)pyrazolylborate (KTp(tBu,Me)). Both Tm(II) compounds were found active as ε-caprolactone polymerisation catalysts.

4.
J Phys Chem A ; 115(29): 8295-301, 2011 Jul 28.
Artículo en Inglés | MEDLINE | ID: mdl-21675778

RESUMEN

An effective methodology to deal with the theoretical treatment on the redox chemistry of divalent organolanthanide complexes is reported and has been tested on two representative substrates, pyridine and CO(2), with two different metals (samarium and thulium). An influence of the ancillary ligands, namely, C(5)Me(5) (Cp*) or (2,3,4,5-tetramethylphospholyl) (Tmp), on the one- or two-electron oxidation processes is observed. The theoretical results are in excellent agreement with the experimental observations indicating the efficiency of the method.

5.
Dalton Trans ; 39(29): 6589-98, 2010 Aug 07.
Artículo en Inglés | MEDLINE | ID: mdl-20631944

RESUMEN

The synthesis of non-classical divalent lanthanide complexes, i.e. those not containing the classical samarium(II), europium(II) or ytterbium(II), was once thought impossible. Since 1997, when the first stable complex of thulium(II) was discovered, there has been many more examples of stable coordination and organometallic complexes of lanthanum(II), neodymium(II) and dysprosium(II) in addition to thulium(II), and the influence of the ligand system on the stability of the complexes is beginning to be understood. These non-classical divalent compounds show exceptional reactivity as some of them have been shown to activate dinitrogen at room temperature, together with related reduced divalent-like systems, and to undergo spontaneous intramolecular carbon-hydrogen bond activation. Many more examples of non-classical divalent compounds together with new aspects of their exciting reactivity should be discovered in the near future.

6.
Dalton Trans ; 39(29): 6761-6, 2010 Aug 07.
Artículo en Inglés | MEDLINE | ID: mdl-20448878

RESUMEN

The bisborohydrido samarium compound Sm(BH(4))(2)(THF)(2) (1) was prepared in high yield by comproportionation between Sm and 2 equivalents of Sm(BH(4))(3)(THF)(3). Reaction of 1 with KCp* (Cp* = C(5)Me(5)) afforded the half-sandwich Cp*Sm(BH(4))(THF)(2) (2). The (1)H NMR borohydride resonances of both complexes, observed at very high field, are typical of divalent Sm compounds. X-Ray single crystal analysis revealed polymeric and dimeric molecular arrangements for 1 and 2, respectively. Complex 1 was found moderately active towards styrene polymerisation when activated with Al((i)Bu)(3), or with a borate/Al combination, whereas 2 showed no activity. Ring-opening polymerisation of epsilon-caprolactone could be performed rapidly at room temperature with both initiators. Complex 2 leads to narrow polydispersities and higher activity. Two different mechanisms, by monoelectronic transfer or by insertion into the Sm-(BH(4)) bond can be proposed.

8.
Dalton Trans ; (42): 4866-70, 2007 Nov 14.
Artículo en Inglés | MEDLINE | ID: mdl-17955139

RESUMEN

Homoleptic complexes Ln(AlMe(4))(3) (Ln = La, Nd) can be straightforwardly utilized in salt metathetic exchange reactions with potassium 2,3,4,5-tetramethylphospholide (KTmp) or 3,4-dimethyl-2,5-bis(trimethylsilyl)phospholide (KDsp) affording monophosphacyclopentadienyl hydrocarbyl complexes (eta(5)-PC(4)Me(4))Ln(AlMe(4))(2) and [eta(5)-PC(4)Me(2)(SiMe(3))(2)]Ln(AlMe(4))(2) (Ln = La, Nd). The solid-state structures reveal distinct metal size effects as evidenced by X-ray diffraction analyses of monomeric neodymium Tmp and Dsp derivatives as well as the dimeric lanthanum Tmp complex. Dimerization is accomplished by intermolecular P --> La donor contacts. Upon activation with [PhMe(2)NH][B(C(6)F(5))(4)] the monophosphacyclopentadienyl complexes initiate the polymerization of isoprene producing 1,4-trans-polyisoprene (tPIP > 87%) with moderate activity (approximately 30 kg(PIP) mol(Ln)(-1) h(-1)).

9.
Chem Commun (Camb) ; (4): 426-8, 2006 Jan 28.
Artículo en Inglés | MEDLINE | ID: mdl-16493824

RESUMEN

The new stable, neutral Tm(II) complex (Cp(ttt))2Tm [Cp(ttt) = 1,2,4-tris(tert-butyl)cyclopentadienyl] can be obtained either by direct reaction of NaCp(ttt) with TmI2 or by reduction of (Cp(ttt))2TmI in non-polar solvents; this latter route may prove itself useful for the isolation of other neutral non-classical low-valent organolanthanide species.

11.
Chemistry ; 9(20): 4916-23, 2003 Oct 17.
Artículo en Inglés | MEDLINE | ID: mdl-14562310

RESUMEN

Potassium 2,5-di-tert-butyl-3,4-dimethylphospholide K(dtp) (9) was synthesised in 45 % yield from commercially available starting materials by using zirconacyclopentadiene chemistry. Reaction of the K salt of this bulky anion and of the previously described potassium 2,5-bis(trimethylsilyl)-3,4-dimethylphospholide K(dsp) (8) with SmI(2) in diethyl ether afforded the homoleptic samarium(II) complexes 7 and 6, respectively, whose solid-state structures, [[Sm(dtp)(2)](2)] (7 a) and [[Sm(dsp)(2)](2)] (6 a), are dimeric owing to coordination of the phosphorus lone pairs to samarium, as shown by X-ray crystallography. Reaction of 8 with TmI(2) in diethyl ether afforded [Tm(dsp)(2)(Et(2)O)], which could not be desolvated without decomposition. In contrast, the coordinated ether group of the solvate [Tm(dtp)(2)(Et(2)O)], obtained from 9 and TmI(2), could easily be removed by evaporation of the solvent and extraction with pentane at room temperature, and the monomer [Tm(dtp)(2)] (5) could be isolated and was characterised by X-ray crystallography. Presumably, steric crowding in 5 is too high for dimerisation to occur. Compound 5, the first Tm(II) homoleptic sandwich complex, is remarkably stable at room temperature in solution and did not noticeably react with nitrogen, in sharp contrast with other thulium(II) species. As expected, 5, 6 and 7 all reacted with azobenzene to give the trivalent complexes [Tm(dtp)(2)(N(2)Ph(2))] (13), [Sm(dsp)(2)(N(2)Ph(2))], (14) and [Sm(dtp)(2)(N(2)Ph(2))] (15), respectively; 13 and 14 were characterised by X-ray crystallography. Complex 5 immediately reacted with triphenylphosphane sulfide at room temperature to give [[Tm(dtp)(2)](2)(mu-S)] (16), which was characterised by X-ray crystallography, whereas samarium(II) complexes 6 and 7 did not noticeably react with Ph(3)PS over 24 h under the same conditions.

12.
Chem Commun (Camb) ; (15): 1646-7, 2002 Aug 07.
Artículo en Inglés | MEDLINE | ID: mdl-12170824

RESUMEN

Reaction of thulium diiodide with substituted phospholide and arsolide salts respectively afforded stable bis(phospholyl)- and bis(arsolyl)thulium(II) complexes, that were characterised by multinuclear NMR and X-ray crystal structures, thus showing the beneficial effects of the steric and electronic properties of crowded phospholyl and arsolyl ligands for the stabilisation of divalent thulium.

SELECCIÓN DE REFERENCIAS
DETALLE DE LA BÚSQUEDA