首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Manganese tricarbonyl complexes (η5-C5H4CH2CH2Br)Mn(CO)3 (3) and (η5-C5H4CH2CH2I)Mn(CO)3 (4), with an alkyl halide side chain attached to the cyclopentadienyl ligand, were synthesized as possible precursors to chelated alkyl halide manganese complexes. Photolysis of 3 or 4 in toluene, hexane or acetone-d6 resulted in CO dissociation and intramolecular coordination of the alkyl halide to manganese to produce (η51-C5H4CH2CH2Br)Mn(CO)2 (5) and (η51-C5H4CH2CH2I)Mn(CO)2 (6). Low temperature NMR and IR spectroscopy established the structures of 5 and 6. Photolysis of 3 in a glass matrix at 91 K demonstrated CO release from manganese. Low temperature NMR spectroscopy established that the coordinated alkyl halide complexes are stable to approximately −20°C.  相似文献   

2.
The kinetics in heptane of displacement of the alkene ligands ethene and methyl acrylate from Ru(CO)42-alkene) by P(OEt)3 have been measured. The reactions occur by reversible dissociation of the alkenes, and activation parameters are compared with those for dissociation of CO from Ru(CO)5 and for reactions of the corresponding Os complexes. A linear free energy relationship for ligand dissociation from Ru(CO)5, Ru(CO)4(C2H4) and Ru(CO)4(MA) has a gradient close to unity, indicating virtually complete bond breaking in the transition states. Competition parameters for reactions of what is probably a solvated Ru(CO)4S intermediate have been measured for the alkenes and P(OEt)3, and for eleven other P-donor nucleophiles. Correlations with the electronic and steric properties of the P-donors show negligible dependence on the electron donicity of the nucleophiles and a small but significant dependence on their sizes. The sizes were quantified by Tolman cone angles or by ‘cone angle equivalents’ derived directly from Brown's ligand repulsion energies (Er). These correlations compared with those, reported elsewhere, for reactions of the probably solvated intermediates Co2(CO)52-C2Ph2) and H3Re3(CO)11 formed by ligand dissociative processes. In all cases the discrimination between nucleophiles by the intermediates is weak confirming their high reactivity and the borderline nature of the mechanisms of these bimolecular reactions between Id and Ia.  相似文献   

3.
The ligand 1,2-bis(3,5-dimethylpyrazol-1-yl)ethane has been synthesized by the direct reaction of 1,2-dibromoethane and 3,5-dimethylpyrazole. The complex of this ligand with palladium(II) chloride has been prepared and the structure of its toluene solvate has been determined by X-ray crystallography. The compound crystallises in the triclinic space group , a = 11.088(5), B = 13.786(5), C = 21.169(9) Å, = 86.96(3), β = 81.02(3), γ = 73.32(3)° with Z = 2. Final residuals after least-squares refinement: R = 0.030, Rw = 0.039. The compound has a trimeric structure which may be described as a ‘molecular tricorn’: the ligand bridges adjacent palladium centres, giving rise to a 21-membered trimetallic macrocycle. The overall structure closely approximates D3 symmetry with approximate two-fold axes passing through each palladium atom and the centre of the 1,2-ethanediyl moiety opposite. An interesting feature of the structure is the close approach of several hydrogen atoms from the 1,2-ethanediyl groups to each palladium centre; these interactions are thought to be close-packing rather than agostic bonds.  相似文献   

4.
A dinuclear Mn(II) di(μ-hydroxo) complex having hydrotris(3,5-diisopropyl-1-pyrazolyl)borate (=TpiPr2) reacted with benzoic acid to yield a dinuclear Mn(II) tri(μ-carboxylato) complex, TpiPr2Mn-(μ-OBz)3-Mn(TpiPr2H). X-ray crystallography reveals the unsymmetrical coordination environments for the manganese centers. One of the two TpiPr2 ligands, which bound to the five-coordinated Mn center, is protonated by the action of the third carboxylic acid and the resulting non-Mn-binding N–H moiety forms an intramolecular hydrogen bond with the oxygen donor of a carboxylate ligand. Steric congestion in the bimetallic core results in the large separation of the manganese centers bridged by the syn-anti carboxylate ligand.  相似文献   

5.
Me3Si-X reagents react to completion at 25°C in a short time to convert Ir(H)2FL2 (L=PtBu2Ph) to Ir(H)2XL2. This involves formation of Ir-O, Ir-N, Ir-I, Ir-S and Ir-C(sp) bonds. Products include some η2-X ligands such as carboxylate and acetamide, NHC(O)CH3. The acetamide is shown to be η2 in the solid state and in solution, but readily rearranges, by a transition state with Ir-O bond cleavage, to effect site exchange of the two inequivalent hydrides. The same synthetic approach succeeds for the more crowded metallated species and these reactions arc shown to fail when F is replaced by Cl in the iridium reagent. Unsaturation at Ir is suggested to be central to the mechanism of these F/X transposition reactions.  相似文献   

6.
The 16-electron complex (CO)4W=C(NMe2)SiPh2Me (1) was photochemically prepared from (CO)5W=C- (NMe2)SiPh2Me. Reactions with selected nucleophiles, having different ligand properties, were performed to test the strength of the intramolecular agostic interaction of one of the phenyl groups, by which 1 is stabilized. The stable complexes cis-(CO)4LW=C(NMe2)SiPh2Me were formed with L=P(OMe)3, P(OEt)3 or 2,6-Me2C6H3NC. The substituted complexes had no tendency for ligand elimination. Addition of acetonitrile or pyridine to an ether solution of 1 resulted in the formation of cis-(CO)4(MeCN)W=C(NMe2)SiPh2Me or cis-(CO)4(C5H5N)- W=C(NMe2)SiPh2Me, respectively. These reactions were reversed on evaporation of the solutions. No reaction was observed with triethylamine.  相似文献   

7.
Treatment of 1-tributylstannylalkynes with Cp2Zr(H)CI affords olefinic intermediates substituted by both Bu3Sn and Cp2ZrCI groups on the terminal sp2-1ike carbon. These stereodefined reagents can be selectively transmetalated at zirconium to afford cuprates which deliver product vinyl stannanes in both substitution and Michael addition reactions.  相似文献   

8.
Selective C-H bond activation of arenes catalyzed by methylrhenium trioxide   总被引:1,自引:0,他引:1  
Arenes, in glacial acetic acid, are oxidized to para-benzoquinones by hydrogen peroxide when methylrhenium trioxide (CH3ReO3 or MTO) is used as a catalyst. In some cases an intermediate hydroquinone was also obtained in lower yield. Oxidation of the methyl side chains of various methylbenzenes did not occur. The active catalyst species are the previously-characterized η2-peroxorhenium complexes, CH3Re(O)2(η2-O2) and CH3Re(O)(η2-O2)2H2O). Separate tests showed that hydroquinones and phenols are oxidized by H2O2-MTO more rapidly than the simple arenes; in the proposed mechanism they are intermediate products. Higher conversions were found for the more highly-substituted arches, consistent with their being the most reactive species toward the electrophillically-active peroxide bound to rhenium. High conversions of the less substituted members of the series were not achieved, reflecting concurrent deactivation of MTO-peroxide, a process of greater import for the more slowly-reacting substrates.  相似文献   

9.
The reactions of Ru(NH3)5py2+, Ru(NH3)4bpy2+, Ru2(NH3)10pz5+, RuRh(NH3)10pz5+ and Ru(NH3)5pz2+ with bromine are first-order in ruthenium and first-order in bromine. The rates decrease with increasing bromide ion concentration and, except for Ru(NH3)5pz2+, are independent of hydrogen ion concentration. The reactions are postulated to proceed via outer-sphere, one-electron transfer from Ru(II) to Br2 with the formation of Br2 as a reactive intermediate. The bromide inhibition is ascribed to the formation of Br3 which is unreactive in outer-sphere reactions because of the barrier imposed by the need to undergo reductive cleavage. The reaction of Ru(NH3)5pz2+ is inhibited by hydrogen ions. The hydrogen ion dependence shows that Ru(NH3)5pzH3+ has a pKa of 2.49 and is at least 500 times less reactive than Ru(NH3)5pz2+. The reaction of Ru2(NH3)10pz4+ with bromine is biphasic. The second phase has a rate identical to that of the Ru2(NH3)10pz5+-Br2 reaction. A detailed analysis shows that the reaction of Ru2(NH3)10pz4+ with bromine proceeds by a sequence of one-electron steps, Br2 being produced as an intermediate. A linear free energy relationship between rate constants and equilibrium constants, obeyed for all the reactions studied, provides an estimate of 1.5 × 102 M−1 s−1 for the self-exchange rate constant of the Br2/Br2 couple.  相似文献   

10.
Treatment of the A-ring aromatic steroids estrone 3-methyl ether and β-estradiol 3, 17-dimethyl ether with Mn(CO)5+BF4 in CH2Cl2 yields the corresponding [(steroid)Mn(CO)3]BF4 salts 1 and 2 as mixtures of and β isomers. The X-ray structure of [(estrone 3-methyl ether)Mn(CO)3]BF4 · CH2Cl2 (1) having the Mn(CO)3 moiety on the side of the steroid is reported: space group P21 with a=10.3958(9), b=10.9020(6), c=12.6848(9) Å, β=111.857(6)°, Z=2, V=1334.3(2) Å3, calc=.481 cm−3, R=0.0508, and wR=0.0635. The molecule has the traditional ‘piano stool’ structure with a planar arene ring and linear Mn---C---O linkages. The nucleophiles NaBH4 and LiCH2C(O)CMe3 add to [(β-estradiol 3,17-dimethyl ether)Mn(CO)3]BF4 (2) in high yield to give the corresponding - and β-cyclohexadienyl manganese tricarbonyl complexes (3). The nucleophiles add meta to the arene -OMe substituent and exo to the metal. The and β isomers of 3 were separated by fractional crystallization and the X-ray structure of the β isomer with an exo-CH2C(O)CMe3 substituent is reported (complex 4): space group P212121 with a=7.5154(8), b=15.160(2), c=25.230(3) Å, Z=4, V=2874.4(5) Å3, calc=1.244 g cm−3, R=0.0529 and wR2=0.1176. The molecule 4 has a planar set of dienyl carbon atoms with the saturated C(1) carbon being 0.592 Å out of the plane away from the metal. The results suggest that the manganese-mediated functionalization of aromatic steroids is a viable synthetic procedure with a range of nucleophiles of varying strengths.  相似文献   

11.
A series of polymer- and silica-supported manganese phosphine complexes has been prepared and characterized. These complexes react reversibly with molecular oxygen in the solid state to yield 1:1 Mn:O2 adducts. The reaction may be reversed either by a pressure drop or a temperature rise. All the O2 adducts are highly colored and binding curves as a function of the partial pressure of oxygen have been constructed. The silica-supported complexes can be prepared from ligand-silanes either by reaction with dehydrated silica and then with anhydrous manganese(II) bromide or vice versa.  相似文献   

12.
Tri-1-alkynyltin compounds [R2Sn(CCR1)3 (1), R2 = Me, R1 = Me (a), nBu (b), tBu (c), Me3Si (d), 1-(1-cyclohexenyl) (e); R2 = Et, R1 = Me (a(Et)), nBu (b(Et)), tBu (c(Et)), SiMe3 (d(Et)); R2 = nBu, R1 = Me (a(Bu)), nBu (b(Bu))] were prepared, and their reactivity towards trialkylboranes Et3B (2) and iPr3B (3) in 1,1-organoboration reactions was studied. The first step in each reaction is an intermolecular 1,1-alkytoboration. Afterwards, intramolecular 1,1-vinyloboration or 1,1-alkyloboration compete with further intermolecular 1,1-alkyloboration. Various triorganotin cations (4-7), stabilized by intramolecular side-on coordination to the CC bond of an alkynylborate moiety, were detected as highly fluxional intermediates prior to rearrangement into heterocyclic systems such as stannoles (9-11), 1-stanna-4-bora-2, 5-cyclohexadienes (8, 12). The reactions between 1a or 1a(Bu) and an excess of Et3B (2) afford the tris(alkenyl)tin compounds 13 via threefold intermolecular 1, 1-ethyloboration. 13 rearrange to the 3-stannolenes (14a or 14a(Bu)). The intermediates and final products were characterized by multinuclear one- and two-dimensional 1H, 11B, 13C, 29Si and 119Sn NMR.  相似文献   

13.
The stability constants of the 1:1 complexes formed between Cu(Arm)2+, where Arm = 2,2′-bipyridyl or 1,10-phenanthroline, and methyl phosphate, CH3OPO32−, or hydrogen phosphate, HOPO32−, were determined by potentiometric pH titration in aqueous solution (25°C; l = 0.1 M, NaNO3). On the basis of previously established log K versus pKa straight-line plots (D. Chen et al., J. Chem. Soc., Dalton Trans. (1993) 1537–1546) for the complexes of simple phosphate monoesters and phosphonate derivatives, R-PO32−, where R is a non-coordinating residue, it is shown that the stabilities of the Cu(Arm) (CH3OPO3) complexes are solely determined by the basicity of the -PO32− residue. In contrast, the Cu(Arm) (HOPO3) complexes are slightly more stable (on average by 0.15 log unit) than expected on the basicity of HPO42−; this is possibly due to a more effective solvation including hydrogen bonding, an interaction not possible with coordinated CH3OPO32− species. Regarding biological systems the observation that HOPO32− is somewhat favored over R-PO32− species in metal ion interactions is meaningful.  相似文献   

14.
The complex [(PP3)OsH(N2)]BPh4 is a catalyst precursor for the regio- and stereoselective dimerization of HCCR (R=Ph, SiMe3) to (Z)-1,4-disubstituted-but-3-en-l-ynes (PP3=P(CH2CH2PPh2)3). In the presence of H2O or C2H5OH, the catalytic reaction with HCCSiMe3 selectively gives but-3-en-l-ynyl-trimethyisilane. A detailed study under different experimental conditions, the detection of some intermediates, and the use of isolated complexes in independent reactions, taken altogether, permit mechanistic conclusions which account for the observed products. A key-role is played by (vinylidene)σ-alkynyl complexes which transform into η3-butenynyl derivatives via intramolecular C---C bond formation. The Os(II) η3-butenynyl complexes are likely reagents in the rate determining step of the catalytic cycle, and produce free (Z)-1,4-disubstituted-but-3-en-l-ynes upon σ-bond metathesis reaction with HCCR. The 16-electron fragments [(PP3)OsX]+ (X = H, Cl, CCR) are capable of promoting the 1-alkyne to vinylidene tautomerism. In particular, the (vinylidene)hydride [(PP3)OsH{C=C(H)-SiMe3}]BPh4 has been isolated and properly characterized. Since the stoichiometric reaction of the latter compound with HCCSiMe3 gives vinyltrimethylsilane, the formation of (vinylidene)hydride species is suggested to be an effective step, alternative to 1-alkyne insertion, in the reduction of 1-alkynes to alkenes assisted by hydrido metal complexes.  相似文献   

15.
16.
Reaction of the allylidene tungsten complex [W(CPhCHCHMe)Br2(CO)2(4-picoline)] (1) with the dithiocarbamates MS2CNR2 (a: M=Na, R=Et; b: M=Na, R=Me; c: M=Li, R=Ph) in THF at 50 °C affords the vinylketene tungsten complexes [W(S2CNR2)2(OCCPhCHCHMe)(CO)] (2a–c). At lower temperatures, four reaction intermediates (3–6) may be discerned. Spectroscopic studies indicate that these compounds contain η4-allyldithiocarbamate ligands which are generated by addition of dithiocarbamate across the metal-carbon double bond of the allylidene-tungsten unit in 1. The structures of [W(S2CNEt2)2(OCCPhCHCHMe)(CO)] (2a) and of one intermediate, [W(η4-Et2NCS2CPhCHCHMe)(S2CNEt2)(CO)2] (5a) were elucidated by X-ray crystallography.  相似文献   

17.
Kinetic and activation parameter data for the reactions of cct-Ru(H)2(CO)2(PPh3)2 (1) (cct = cis, cis, trans) in THF with thiols, CO and PPh3 to give cct-RuH(SR)(CO)2(PPh3)2, Ru(CO)3(PPh3)2 and Ru(CO)2(PPh3)2, respectively, reveal a common, rate-determining step, the initial dissociation of H2 from 1; the activated complex probably resembles the corresponding Ru(η2-H2) species. Reaction of Ru(H)2(dppm)2 (2) (as a cis/trans mixture, DPPM = bis(diphenylphosphino)methane) with thiols initially generated cis- and trans- RuH(SR) (dppm)2 with a rate that depends on both the type and concentration of thiol. The higher basicity of the hydride ligands in 2 (versus 1), which is demonstrated by deuterium exchange with CD3OD, gives rise in the thiol reaction to an initial protonation step prior to loss of H2. A species detected in the thiol reaction is possibly [RuH(η2-H2 (dppm)2]2, the anticipated intermediate for this reaction and for the hydrogen exchange with alcohol. A longer reaction of 2 with PhCH2SH gives solely cis-Ru(SCH2Ph)2(dppm)2.  相似文献   

18.
Tricarbonyl(η6-1-oxobenzocyclobutene)chromium(0) (1) can be transformed to tricarbonyl(η6-1-endo-hydroxybenzocyclobutene) chromium(0) derivatives with substituents R (R=CH3, CH=CH2, (CH2)4CH=CH2, (CH2)4OSi(Me)2tBu) at Cl on the exo face of the complex. The relative configuration is proven by an X-ray crystal structure analysis of the trimethylsilyl ether 8 (C16H18CrO4Si: a=8.693(1), b=9.490(1), c=11.063(1) Å, =97.51(1), β=110.32(1), γ=95.38(1)°, triclinic, space group P (No.2), R=0.037, Rw=0.052 for 4609 observed reflections. Attempts directed at an intramolecular cycloaddition of the ortho-quinodimethane complex derived from 17 by anion promoted ring opening unexpectedly resulted in the formation of 12 as the product of an opening of the proximal bond of the anellated ring located between the hydroxy group and the coordinated aromatic ring in 16. The fact that the intermolecular cycloaddition reaction for 16 is possible in the presence of a dienophile is taken as evidence for an equilibrium between the alcoholate 17 and the two ring opened products 16 and 18. The proximal ring opening of 6 is not observed when the free organic ligand 21 is used as the educt. Ketone complexes 1 and 25 undergo proximal ring opening reaction when treated with alcoholate or primary amines.  相似文献   

19.
The mononuclear manganese(III) complexes [C5H10NH2][MnL2] [L2−=a substituted N-(2-hydroxybenzyl)glycinate (hbg2−) viz. 3,5-dibromo- (3,5-Br-hbg2−), 3,5-dichloro- (3,5-Cl-hbg2−), 3-methyl-5-chloro- (3,5-Me,Cl-hbg2−), 5-bromo- (5-Br-hbg2−), 5-chloro- (5-Cl-hbg2−), 5-nitro- (5-NO2-hbg2−) or N-(5-nitro-2-hydroxybenzyl)sarcosine (5-NO2-hbs2−)] have been synthesised by reaction of the appropriate ligand with manganese(II) perchlorate under ambient conditions in a 2:1 molar ratio using piperidine as base. The structures of three of these complexes, [C5H10NH2][Mn(3,5-Cl-hbg)2] (2), [C5H10NH2][Mn(5-NO2-hbg)2] (6) and [C5H10NH2][Mn(5-NO2-hbs)2] (7) have been elucidated by single-crystal X-ray crystallography and each displays two similar, independent [MnL2] ions in the asymmetric unit linked via piperidinium cations through hydrogen bonding. The ligands co-ordinate in a facial tridentate fashion with the three donor atoms being the phenolate and carboxylate oxygens and the amine nitrogen. The geometry at the Mn centres is compressed rhombic octahedral consistent with a pseudo-Jahn–Teller compression along the Mn–O(phenolate) axis. Mean bond lengths are in the ranges 1.886–1.889 Å for the Mn–O(phenolate), 2.062–2.125 Å for the Mn–O(carboxylate) and 2.091–2.184 Å for the Mn–N(amine) distances. The magnetic susceptibility and electronic and IR spectroscopic data are discussed with reference to the crystal structures.  相似文献   

20.
Copper(I)/(II) complexes with the ligand 2-aminoethyl(2-pyridylmethyl)1,2-ethanediamine (apme, abbreviated as PDT in the literature as well) were prepared and characterized. Crystal structures of the copper(I) complexes, [Cu2(apme)2]X2 (1, 2; X = ClO4, CF3SO3), showed that they are dinuclear, in contrast to the trigonal bipyramidal copper(II) complexes [Cu(apme)Cl]BPh4 (3) and [Cu(apme)(DMF)](BPh4)2 (4). 1 and 2 could be investigated in solution by NMR spectroscopy and 3 and 4 by cyclovoltammetry. From the results of these studies it is clear that in solution equilibria between the dinuclear complexes 1/2 and another species exist, most likely the monomeric [Cu(apme)CH3CN]+. Time-resolved UV/vis spectra at low temperatures allowed the spectroscopic detection of dioxygen adduct complexes as reactive intermediates during the oxidation of 1/2 with dioxygen that seem to play an important role in copper enzymes such as peptidylglycine--hydroxylating monooxygenase (PHM).  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号