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











Base de datos
Intervalo de año de publicación
1.
Organometallics ; 31(7): 2631-2638, 2012.
Artículo en Inglés | MEDLINE | ID: mdl-23087535

RESUMEN

The bis(trimethylsilyl)-substituted hydroxycyclopentadienyl ruthenium hydride [2,5-(SiMe(3))(2)-3,4-(CH(2)OCH(2))(η(5)-C(4)COH)]Ru(CO)(2)H (10) is an efficient catalyst for hydrogenation of aldehydes and ketones. Because 10 transfers hydrogen rapidly to aldehydes and ketones and because it does not form an inactive bridging hydride during reaction, hydrogenation of aldehydes and ketones can be performed at room temperature under relatively low hydrogen pressure (3 atm); this is a significant improvement compared with previously developed Shvo type catalysts. Kinetic and (2)H NMR spectroscopic studies of the stoichiometric reduction of aldehydes and ketones by 10 established a two-step process for the hydrogen transfer: (1) rapid and reversible hydrogen bond formation between OH of 10 and the oxygen of the aldehyde or ketone, (2) followed by slow transfer of both proton and hydride from 10 to the aldehyde or ketone. The stoichiometric and catalytic activities of complex 10 are compared to those of other Shvo type ruthenium hydrides and related iron hydrides.

2.
Chem Commun (Camb) ; 46(42): 7915-7, 2010 Nov 14.
Artículo en Inglés | MEDLINE | ID: mdl-20856982

RESUMEN

The ruthenium hydride of (Ar(4)CpOH)Ru(CO)(2)H exchanges cleanly and rapidly with D(2) at room temperature to generate the ruthenium deuteride. A chain mechanism is proposed to explain the much more rapid exchange of RuH/D(2) than RuCO exchange with (13)CO.


Asunto(s)
Deuterio/química , Rutenio/química , Enlace de Hidrógeno , Espectroscopía de Resonancia Magnética
3.
J Am Chem Soc ; 131(7): 2499-507, 2009 Feb 25.
Artículo en Inglés | MEDLINE | ID: mdl-19193034

RESUMEN

Cyclopentadienone iron alcohol complexes generated from the reactions of [2,5-(SiMe(3))(2)-3,4-(CH(2))(4)(eta(5)-C(4)COH)]Fe(CO)(2)H (3) and aldehydes were characterized by (1)H NMR, (13)C NMR, and IR spectroscopy. The benzyl alcohol complex [2,5-(SiMe(3))(2)-3,4-(CH(2))(4)(eta(4)-C(4)CO)]Fe(CO)(2)(HOCH(2)C(6)H(5)) (6-H) was also characterized by X-ray crystallography. These alcohol complexes are thermally unstable and prone to dissociate the coordinated alcohols. The alcohol ligand is easily replaced by other ligands such as PhCN, pyridine, and PPh(3). Dissociation of the alcohol ligand in the presence of H(2) leads to the formation of iron hydride 3. The reduction of aldehydes by 3 was carried out in the presence of both potential intermolecular and intramolecular traps. The reaction of 3 with PhCHO in the presence of 4-methylbenzyl alcohol as a potential intermolecular trapping agent initially produced only iron complex 6-H of the newly formed benzyl alcohol. However, the reaction of 3 with 4-(HOCD(2))C(6)H(4)CHO, which possesses a potential intramolecular alcohol trapping agent, afforded two alcohol complexes, one with the newly formed alcohol coordinated to iron and the other with the trapping alcohol coordinated. The intramolecular trapping experiments support a mechanism involving concerted transfer of a proton from OH and hydride from Fe of 3 to aldehydes. The kinetics and mechanism of the hydrogenation of benzaldehyde catalyzed by 3 are presented.


Asunto(s)
Alcoholes/química , Aldehídos/química , Compuestos de Hierro/química , Alcoholes/síntesis química , Benzaldehídos/química , Alcohol Bencilo/química , Catálisis , Cristalografía por Rayos X , Hidrogenación , Compuestos de Hierro/síntesis química , Cinética , Espectroscopía de Resonancia Magnética/métodos , Nitrilos/química , Compuestos Organofosforados/química , Oxidación-Reducción , Piridinas/química , Espectrofotometría Infrarroja , Termodinámica
4.
Organometallics ; 28(1): 123-131, 2009.
Artículo en Inglés | MEDLINE | ID: mdl-20046997

RESUMEN

PMe(3) adds selectively to the central carbon of the η(3)- propargyl complex [C(5)Me(5)(CO)(2)Re(η(3)-CH(2)C≡CCMe(3))][BF(4)] (1-t-Bu) to form the metallacyclobutene [C(5)Me(5)(CO)(2)Re(CH(2)C(PMe(3))=CCMe(3))][BF(4)] (7). The rate of rearrangement of the metallacyclobutene 7 to η(2)-alkyne complex [C(5)Me(5)(CO)(2)Re(η(2)-Me(3)PCH(2)C≡CCMe(3))][BF(4)] (8) is is independent of phosphine concentration, consistent with a dissociative mechanism proceeding via η(3)-propargyl complex 1-t-Bu. The rate of this rearrangement is 480 times slower than the rate of exchange of PMe(3) with the labeled metallacyclobutene 7-d(9). This rate ratio provides an indirect measurement of the regioselectivity for addition of PMe(3) to the central carbon of η(3)-propargyl complex 1-t-Bu to give 7 compared to addition to a terminal carbon to give 8. The addition of PPh(3) to 1-t-Bu gives the metallacyclobutene [C(5)Me(5)(CO)(2)Re(CH(2)C(PPh(3))=CCMe(3))][BF(4)] (11). Low temperature (1)H NMR spectra provide evidence for an equilibrium between metallacyclobutene 11 and η(3)-propargyl complex 1-t-Bu (K(eq) ≈ 44 M(-1) at -46 °C and ΔG° (0 °C) = -1.2 ± 0.2 kcal mol(-1)).

5.
J Am Chem Soc ; 130(7): 2285-95, 2008 Feb 20.
Artículo en Inglés | MEDLINE | ID: mdl-18215043

RESUMEN

The catalytic hydrogenation of benzaldehyde and acetophenone with the Shvo hydrogenation catalysts were monitored by in situ IR spectroscopy in both toluene and THF. The disappearance of organic carbonyl compound and the concentrations of the ruthenium species present throughout the hydrogenation reaction were observed. The dependence of the hydrogenation rate on substrate, H2 pressure, total ruthenium concentration, and solvent were measured. In toluene, bridging diruthenium hydride 1 was the only observable ruthenium species until nearly all of the substrate was consumed. In THF, both 1 and some monoruthenium hydride 2 were observed during the course of the hydrogenation. A full kinetic model of the hydrogenation based on rate constants for individual steps in the catalysis was developed. This kinetic model simulates the rate of carbonyl compound hydrogenation and of the amounts of ruthenium species 1 and 2 present during hydrogenations.


Asunto(s)
Acetofenonas/química , Benzaldehídos/química , Hidrógeno/química , Compuestos Organometálicos/química , Catálisis , Simulación por Computador , Hidrogenación , Cinética , Modelos Químicos , Rutenio/química , Espectrofotometría Infrarroja , Tolueno/química
6.
J Am Chem Soc ; 129(38): 11821-7, 2007 Sep 26.
Artículo en Inglés | MEDLINE | ID: mdl-17760443

RESUMEN

Reduction of imines by [2,5-Ph2-3,4-Tol(2)(eta(5)-C(4)COH)]Ru(CO)2H (1) produces kinetically stable ruthenium amine complexes. Reduction of an imine possessing an intramolecular amine was studied to distinguish between inner sphere and outer sphere mechanisms. 1,4-Bn(15)NH(c-C(6)H(10))=NBn (12) was reduced by 1 in toluene-d8 to give 85% of [2,5-Ph2-3,4-Tol(2)(eta(4)-C(4)CO)](CO)(2)RuNHBn(c-C(6)H(10))(15)NHBn (16-RuN,15N), resulting from coordination of the newly formed amine to the ruthenium center, and 15% of trapping product [2,5-Ph2-3,4-Tol(2)(eta(4)-C(4)CO)](CO)(2)Ru(15)NHBn(c-C(6)H(10))NHBn (16-Ru(15)N,N), resulting from coordination of the intramolecular trapping amine. These results provide support for an outer sphere transfer of hydrogen to the imine to generate a coordinatively unsaturated intermediate, which can be trapped by the intramolecular amine. An opposing mechanism, requiring coordination of the imine nitrogen to ruthenium prior to hydrogen transfer, cannot readily explain the observation of the trapping product 16-Ru(15)N,N.


Asunto(s)
Iminas/química , Compuestos Organometálicos/química , Aminas/química , Hidrógeno/química , Cinética , Conformación Molecular , Oxidación-Reducción , Rutenio/química , Estereoisomerismo , Tolueno/química
9.
J Am Chem Soc ; 128(7): 2286-93, 2006 Feb 22.
Artículo en Inglés | MEDLINE | ID: mdl-16478183

RESUMEN

The stereochemistry of hydrogen transfer from [2,5-Ph(2)-3,4-Tol(2)(eta(5)-C(4)COD)]Ru(CO)(2)D to N-aryl imines to give amine complexes was shown to be mostly trans stereospecific. Stereospecific hydrogen transfer is proposed to generate an amine and a coordinatively unsaturated ruthenium intermediate in close proximity. Coordination of the amine is proposed to occur faster than lone pair inversion of the amine. In contrast, hydrogen transfer to N-alkyl imines is stereorandom. It is proposed that stereochemistry is lost in part due to the reversibility of the hydrogen transfer being faster than amine coordination.


Asunto(s)
Iminas/química , Compuestos Organometálicos/química , Medición de Intercambio de Deuterio , Espectroscopía de Resonancia Magnética , Conformación Molecular , Oxidación-Reducción , Estereoisomerismo , Termodinámica
10.
J Am Chem Soc ; 127(40): 14062-71, 2005 Oct 12.
Artículo en Inglés | MEDLINE | ID: mdl-16201828

RESUMEN

Reduction of imines by [2,5-Ph2-3,4-Tol2(eta(5)-C4COH)]Ru(CO)2H (2) produces kinetically stable ruthenium amine complexes. Reduction of an imine by 2 in the presence of an external amine trap gives only the complex of the newly generated amine. Reaction of 2 with H2N-p-C6H4N=CHPh (11), which contains an intramolecular amine trap, gave a 1:1 mixture of [2,5-Ph2-3,4-Tol2(eta(4)-C4CO)](CO)2RuNH(CH2Ph)(C6H4-p-NH2) (8), formed by coordination of the newly generated amine to the ruthenium center, and [2,5-Ph2-3,4-Tol2(eta(4)-C4CO)](CO)2RuNH2C6H4-p-NHCH2Ph (9), formed by coordination of the amine already present in the substrate. These results require transfer of hydrogen to the imine outside the coordination sphere of the metal to give a coordinatively unsaturated intermediate that can be trapped inside the initial solvent cage. Amine diffusion from the solvent cage must be much slower than coordination to the metal center. Mechanisms requiring prior coordination of the substrate to ruthenium would have led only to 8 and can be eliminated.


Asunto(s)
Iminas/química , Compuestos Organometálicos/química , Rutenio/química , Cristalografía por Rayos X , Enlace de Hidrógeno , Cinética , Modelos Moleculares , Conformación Molecular , Compuestos Organometálicos/síntesis química , Oxidación-Reducción , Transición de Fase , Protones
11.
J Am Chem Soc ; 127(9): 3100-9, 2005 Mar 09.
Artículo en Inglés | MEDLINE | ID: mdl-15740149

RESUMEN

At high temperatures in toluene, [2,5-Ph(2)-3,4-Tol(2)(eta(5)-C(4)COH)]Ru(CO)(2)H (3) undergoes hydrogen elimination in the presence of PPh(3) to produce the ruthenium phosphine complex [2,5-Ph(2)-3,4-Tol(2)-(eta(4)-C(4)CO)]Ru(PPh(3))(CO)(2) (6). In the absence of alcohols, the lack of RuH/OD exchange, a rate law first order in Ru and zero order in phosphine, and kinetic deuterium isotope effects all point to a mechanism involving irreversible formation of a transient dihydrogen ruthenium complex B, loss of H(2) to give unsaturated ruthenium complex A, and trapping by PPh(3) to give 6. DFT calculations showed that a mechanism involving direct transfer of a hydrogen from the CpOH group to form B had too high a barrier to be considered. DFT calculations also indicated that an alcohol or the CpOH group of 3 could provide a low energy pathway for formation of B. PGSE NMR measurements established that 3 is a hydrogen-bonded dimer in toluene, and the first-order kinetics indicate that two molecules of 3 are also involved in the transition state for hydrogen transfer to form B, which is the rate-limiting step. In the presence of ethanol, hydrogen loss from 3 is accelerated and RuD/OH exchange occurs 250 times faster than in its absence. Calculations indicate that the transition state for dihydrogen complex formation involves an ethanol bridge between the acidic CpOH and hydridic RuH of 3; the alcohol facilitates proton transfer and accelerates the reversible formation of dihydrogen complex B. In the presence of EtOH, the rate-limiting step shifts to the loss of hydrogen from B.


Asunto(s)
Hidrógeno/química , Compuestos Organometálicos/química , Rutenio/química , Catálisis , Medición de Intercambio de Deuterio , Etanol/química , Hidrogenación , Cinética , Modelos Químicos , Modelos Moleculares , Fosfinas/química , Tolueno/química , Agua/química
12.
J Org Chem ; 70(7): 2576-81, 2005 Apr 01.
Artículo en Inglés | MEDLINE | ID: mdl-15787545

RESUMEN

[reactions: see text] The cis-2-alken-4-yn-1-one, 1-phenyl-cis-2-penten-4-yn-1-one (cis-1), readily dimerizes on treatment with weak acid to give the 1,2-difurylethylenes, trans- and cis-1,2 di(2-(5-phenylfuryl))ethene (trans-1 and cis-2), in 62% and 23% yields, respectively. Trimerization of cis-1 to trans,trans-1,2,3-tri(2-(5-phenylfuryl)cyclopropane (4) occurred as a byproduct of treatment with weak acid. These reactions demonstrate the 2-furylcarbenoid reactivity of cis-2-alken-4-yn-1-ones.

13.
J Am Chem Soc ; 127(6): 1883-94, 2005 Feb 16.
Artículo en Inglés | MEDLINE | ID: mdl-15701023

RESUMEN

Hydroxycyclopentadienyl ruthenium hydride 5 efficiently reduces imines below room temperature. Better donor substituents on nitrogen give rise to faster rates and a shift of the rate-determining step from hydrogen transfer to amine coordination. Reduction of electron-deficient N-benzilidenepentafluoroaniline (8) at 11 degrees C resulted in free amine and kinetic isotope effects of k(OH)/k(OD) = 1.61 +/- 0.08, k(RuH)/k(RuD) = 2.05 +/- 0.08, and k(RuHOH)/k(RuDOD) = 3.32 +/- 0.14, indicative of rate-limiting concerted hydrogen transfer, a mechanism analogous to that proposed for aldehyde and ketone reduction. Reduction of electron-rich N-alkyl-substituted imine, N-isopropyl-(4-methyl)benzilidene amine (9), was accompanied by facile imine isomerization and scrambling of deuterium labels from reduction with 5-RuDOH into the N-alkyl substituent of both the amine complex and into the recovered imine. Inverse equilibrium isotope effects were observed in the reduction of N-benzilidene-tert-butylamine (11) at -48 degrees C (k(OH)/k(OD) = 0.89 +/- 0.06, k(RuH)/k(RuD) = 0.64 +/- 0.05, and k(RuHOH)/k(RuDOD) = 0.56 +/- 0.05). These results are consistent with a mechanism involving reversible hydrogen transfer followed by rate-limiting amine coordination.


Asunto(s)
Ciclopentanos/química , Iminas/química , Compuestos Organometálicos/química , Rutenio/química , Compuestos de Anilina/química , Catálisis , Medición de Intercambio de Deuterio , Hidrogenación , Isomerismo , Cinética
14.
Beilstein J Org Chem ; 1(1): 18, 2005 Dec 09.
Artículo en Inglés | MEDLINE | ID: mdl-16542028

RESUMEN

The reaction of di-(o-acetylphenyl)acetylene (1) with excess dimethyl acetylenedicarboxylate (DMAD) produced bis-DMAD adducts meso-3 and rac-3. This transformation is suggested to involve thermal rearrangement of 1 to the intermediate 3,3'-dimethyl-1,1'-biisobenzofuran (A), and subsequent Diels-Alder cycloadditions of two equivalents of DMAD to A. The isolation of trapping products meso-3 and rac-3, which contain complex polycyclic frameworks, provide strong evidence for the transient production of A, the first biisobenzofuran. An X-ray crystal structure of meso-3 was obtained.

15.
J Am Chem Soc ; 126(17): 5585-92, 2004 May 05.
Artículo en Inglés | MEDLINE | ID: mdl-15113230

RESUMEN

Deuterioformylation of styrene catalyzed by [(2S,4S)-BDPP]Pt(SnCl(3))Cl at 39 degrees C gave 3-phenylpropanal (3) and 2-phenylpropanal (2) (n:i = 1.8, 71% ee (S)-2) with deuterium only beta to the aldehyde carbonyl and in the formyl group. Small amounts of deuterium were also found in the internal (2.8%), cis terminal (1.4%), and trans terminal (1.3%) vinyl positions of the recovered styrene. Deuterioformylation of styrene at 98 degrees C gave 3- (3) and 2-phenylpropanal (2) (n:i = 2.3, 10% ee (R)-2) with deuterium both alpha and beta to the aldehyde carbonyl and in the formyl group. Deuterium was also found in the internal (20%), cis terminal (12%), and trans terminal (12%) vinyl positions of the recovered styrene. These deuterioformylation results establish that platinum hydride addition to styrene is largely irreversible at 39 degrees C but reversible at 98 degrees C. Hydroformylation of (E)- and (Z)-beta-deuteriostyrene at 40 degrees C, followed by oxidation of the aldehydes to acids, and subsequent derivitization to the (S)-mandelate esters confirmed that 84% of 2-phenylpropanal (2) arises from platinum hydride addition to the si-face of styrene, while 73% of 3-phenylpropanal (3) arises from platinum hydride addition to the re-face of styrene. At 100 degrees C, the effect of variable H(2) and CO pressure on n:i, % ee, and TOF of hydroformylation of styrene was investigated. The results are consistent with enantioselectivity not being fully determined until the final hydrogenolysis of a platinum acyl intermediate.


Asunto(s)
Compuestos Organoplatinos/química , Compuestos Orgánicos de Estaño/química , Estireno/química , Temperatura , Aldehídos/química , Monóxido de Carbono/química , Catálisis , Deuterio/química , Hidrógeno/química , Espectroscopía de Resonancia Magnética , Estructura Molecular , Presión , Propanoles , Estereoisomerismo
16.
J Am Chem Soc ; 126(6): 1699-704, 2004 Feb 18.
Artículo en Inglés | MEDLINE | ID: mdl-14871100

RESUMEN

The reaction of (Z)-HDC=CHCH(OCH(3))C(6)H(5) (1) with Cp(2)Zr(D)Cl followed by BF(3).OEt(2) gave phenylcyclopropanes 3a and 3b, both having cis deuterium. This stereochemical outcome requires inversion of configuration at the carbon bound to zirconium and is consistent with a "W-shaped" transition state structure for cyclopropane formation. In a Kulinkovich hydroxycyclopropanation, trans-3-deutero-1-methyl-cis-2-phenyl-1-cyclopropanol (5) was formed stereospecifically from Ti(O-i-Pr)(4), ethyl acetate, EtMgBr, and trans-beta-deuterostyrene. This stereochemistry requires retention of configuration at the carbon bound to titanium and is consistent with frontside attack of the carbon-titanium bond on a carbonyl group coordinated to titanium. In a de Meijere cyclopropylamine synthesis, a 3:1 mixture of N,N-dimethyl-N-(trans-3-deutero-trans-2-phenylcyclopropyl)amine (6a) and N,N-dimethyl-N-(cis-3-deutero-cis-2-phenylcyclopropyl)amine (6b) was formed from Ti(O-i-Pr)(4), DMF, Grignard reagents, and trans-beta-deuterostyrene. This stereochemistry requires inversion of configuration at the carbon bound to titanium and is consistent with a W-shaped transition structure for ring closure.

17.
J Org Chem ; 68(5): 1998-2001, 2003 Mar 07.
Artículo en Inglés | MEDLINE | ID: mdl-12608822

RESUMEN

The isotope effects in the reaction of [p-(Me2CH)C6H4Me]Ru(NHCHPhCHPhNSO2C6H4-p-CH3) (1) with isopropyl alcohol were 1.79 for transfer of hydrogen from OH to N and 2.86 for transfer from CH to Ru. The isotope effect for transfer of deuterium from doubly labeled material (kCHOH/kCDOD = 4.88) was within experimental error of the product of the two individual isotope effects. These isotope effects provide convincing evidence for a mechanism involving concurrent transfer of hydrogen from oxygen to nitrogen and from carbon to ruthenium.


Asunto(s)
Compuestos Organometálicos/química , Rutenio/química , 2-Propanol/química , Catálisis , Deuterio/química , Hidrógeno/química , Hidrogenación , Isótopos , Cinética , Modelos Químicos , Estructura Molecular , Oxidación-Reducción
18.
J Am Chem Soc ; 125(9): 2641-51, 2003 Mar 05.
Artículo en Inglés | MEDLINE | ID: mdl-12603152

RESUMEN

Agostic interactions in yttrium alkyls are structure dependent. Primary alkyl yttrium complexes have beta-CH(2) agostic interactions at low temperature, but a shift toward alpha-agostic interactions occurs on warming. For the more crowded beta-disubstituted yttrium alkyls, an alpha-CH(2) agostic interaction is seen. The thermodynamics of alkene binding to the primary alkyl yttrium complex Cp(2)YCH(2)CH(2)CH(CH(3))(2) (2) depend strongly on the structure of the alkene. A single allylic substituent on the alkene has a small effect on alkene binding, but a second allylic substituent has a large destabilizing effect. Propene binding to yttrium alkyls is largely independent of the nature of the alkyl ligand. Equilibrium constants for propene binding to n-, gamma-substituted, beta-substituted, and secondary alkyl yttrium complexes are similar. The rate of migration of an alkyl group to a coordinated alkene depends strongly on the structure of the alkyl group: n-alkyl approximately gamma-substituted >> beta-substituted >> alpha-substituted. The approximately 200-fold slower insertion of propene into Cp(2)YCH(2)CH(CH(3))(2) (6) than that into Cp(2)YCH(2)CH(2)CH(CH(3))(2) (2) is therefore due to kinetically slow migration of the beta-disubstituted alkyl group of 6 and not to differences in the equilibrium binding of propene. Processes related to chain transfer and site epimerization at the metal center are also reported.

19.
J Am Chem Soc ; 124(44): 13215-21, 2002 Nov 06.
Artículo en Inglés | MEDLINE | ID: mdl-12405850

RESUMEN

Hydride abstraction from C(5)Me(5)(CO)(2)Re(eta(2)-PhC triple bond CCH(2)Ph) (1) gave a 3:1 mixture of eta(3)-propargyl complex [C(5)Me(5)(CO)(2)Re(eta(3)-PhCH-C triple bond CPh)][BF(4)] (5) and eta(2)-1-metalla(methylene)cyclopropene complex [C(5)Me(5)(CO)(2)Re(eta(2)-PhC-C=CHPh)][BF(4)] (6). Observation of the eta(2)-isomer requires 1,3-diaryl substitution and is favored by electron-donating substituents on the C(3)-aryl ring. Interconversion of eta(3)-propargyl and eta(2)-1-metalla(methylene)cyclopropene complexes is very rapid and results in coalescence of Cp (1)H NMR resonances at about -50 degrees C. Protonation of the alkynyl carbene complex C(5)Me(5)(CO)(2)Re=C(Ph)C triple bond CPh (22) gave a third isomer, the eta(3)-benzyl complex [C(5)Me(5)(CO)(2)Re[eta(3)(alpha,1,2)-endo,syn-C(6)H(5)CH(C triple bond CC(6)H(5))]][BF(4)] (23) along with small amounts of the isomeric complexes 5 and 6. While 5 and 6 are in rapid equilibrium, there is no equilibration of the eta(3)-benzyl isomer 23 with 5 and 6.

20.
J Am Chem Soc ; 124(11): 2584-94, 2002 Mar 20.
Artículo en Inglés | MEDLINE | ID: mdl-11890809

RESUMEN

Dimerization of the alkynylcarbene complex Cp(CO)(2)Re=C(Tol)C(triple bond)CCH(3) (8) occurs at 100 degrees C to give a 1.2:1 mixture of enediyne complexes [Cp(CO)(2)Re](2)[eta(2),eta(2)-TolC(triple bond)CC(CH(3))=C(CH(3))C(triple bond)CTol] (10-Eand 10-Z), showing no intrinsic bias toward trans-enediyne complexes. The cyclopropyl-substituted alkynylcarbene complex Cp(CO)(2)Re=C(Tol)C(triple bond)CC(3)H(5) (11) dimerizes at 120 degrees C to give a 5:1 ratio of enediyne complexes [Cp(CO)(2)Re](2)[eta(2),eta(2)-TolC(triple bond)C(C(3)H(5))C=C(C(3)H(5))C(triple bond)CTol] (12-E and 12-Z); no ring expansion product was observed. This suggests that if intermediate A formed by a [1,1.5] Re shift and having carbene character at the remote alkynyl carbon is involved, then interaction of the neighboring Re with the carbene center greatly diminishes the carbene character as compared with that of free cyclopropyl carbenes. The tethered bis-(alkynylcarbene) complex Cp(CO)(2)Re=C(Tol)C(triple bond)CCH(2)CH(2)CH(2)C(triple bond)CC(Tol)= Re(CO)(2)Cp (13) dimerizes rapidly at 12 degrees C to give the cyclic cis-enediyne complex [Cp(CO)(2)Re](2)[eta(2),eta(2)-TolC(triple bond)CC(CH(2)CH(2)CH(2))=CC(triple bond)CTol] (15). Attempted synthesis of the 1,8-disubstituted naphthalene derivative 1,8-[Cp(CO)(2)Re=C(Tol)C(triple bond)C](2)C(10)H(6) (16), in which the alkynylcarbene units are constrained to a parallel geometry, leads to dimerization to [Cp(CO)(2)Re](2)(eta(2),eta(2)-1,2-(tolylethynyl)acenaphthylene] (17). The very rapid dimerizations of both 13 and 16 provide compelling evidence against mechanisms involving cyclopropene intermediates. A mechanism is proposed which involves rate-determining addition of the carbene center of A to the remote alkynyl carbon of a second alkynylcarbene complex to generate vinyl carbene intermediate C, and rearrangement of C to the enediyne complex by a [1,1.5] Re shift.

SELECCIÓN DE REFERENCIAS
DETALLE DE LA BÚSQUEDA