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 47 毫秒
1.
[{Rh(μ‐Cl)(H)2(IPr)}2] (IPr = 1,3‐bis‐(2,6‐diisopropylphenyl)imidazole‐2‐ylidene) was found to be an efficient catalyst for the synthesis of novel propargylamines by a one‐pot three‐component reaction between primary arylamines, aliphatic aldehydes, and triisopropylsilylacetylene. This methodology offers an efficient synthetic pathway for the preparation of secondary propargylamines derived from aliphatic aldehydes. The reactivity of [{Rh(μ‐Cl)(H)2(IPr)}2] with amines and aldehydes was studied, leading to the identification of complexes [RhCl(CO)IPr(MesNH2)] (MesNH2 = 2,4,6‐trimethylaniline) and [RhCl(CO)2IPr]. The latter shows a very low catalytic activity while the former brought about reaction rates similar to those obtained with [{Rh(μ‐Cl)(H)2(IPr)}2]. Besides, complex [RhCl(CO)IPr(MesNH2)] reacts with an excess of amine and aldehyde to give [RhCl(CO)IPr{MesN?CHCH2CH(CH3)2}], which was postulated as the active species. A mechanism that clarifies the scarcely studied catalytic cycle of A3‐coupling reactions is proposed based on reactivity studies and DFT calculations.  相似文献   

2.
The reaction of [{Ir(cod)(μ‐Cl)}2] and K2CO3 or of [{Ir(cod)(μ‐OMe)}2] alone with the non‐natural tetrapyrrole 2,2′‐bidipyrrin (H2BDP) yields, depending on the stoichiometry, the mononuclear complex [Ir(cod)(HBDP)] or the homodinuclear complex [{Ir(cod)}2(BDP)]. Both complexes react readily with carbon monoxide to yield the species [Ir(CO)2(HBDP)] and [{Ir(CO)2}2(BDP)], respectively. The results from NMR spectroscopy and X‐ray diffraction reveal different conformations for the tetrapyrrolic ligand in both complexes. The reaction of [{Ir(coe)2(μ‐Cl)}2] with H2BDP proceeds differently and yields the macrocyclic [4e?,2H+]‐oxidized product [IrCl2(9‐Meic)] (9‐Meic = monoanion of 9‐methyl‐9,10‐isocorrole), which can be addressed as an iridium analog of cobalamin.  相似文献   

3.
The reaction of 2,2′-bidipyrrins H2BDP with the RhI complexes [Rh(COD)(μ-OMe)]2 and [Rh(CO)2(μ-Cl)]2 yields the neutral species [{Rh(COD)}2BDP] (7, 8) and [{Rh(CO)2}2BDP] (2, 9), respectively. Treatment of the COD complexes with carbon monoxide results in the exchange of all COD ligands against CO. Ligand exchange studies on the carbonyl complexes 2 and 9 with different phosphane donors reveal the regioselective exchange of one CO per metal ion. In most cases, the reaction is accompanied by a large conformational change of the tetrapyrrole from a syn to an anti conformation. This conformational change was resolved by a combination of NMR spectroscopy and X-ray diffraction studies.  相似文献   

4.
The crystal structure of MoOs2(CO)11[P(OMe)3]2·[(MeO)3P](OC)4OsMo(CO)5 is comprised of a slightly disordered, triangular cluster with a Mo(CO)5 and two Os(CO)3[P(OMe)3] moieties (OsMo bond lengths are 3.041(2) and 3.079(2) Å) together with a [(MeO)3P](OC)4OsMo(CO)5 molecule having a donor-acceptor OsMo bond of length 3.075(2) Å.  相似文献   

5.
Summary The preparations and characterisation of cationic complexes of the type [Rh(CO)(MeCN)(PR3)2]ClO4, [Rh(CO)L(PR3)2]ClO4 (L=py or 2-MeOpy), [Rh(CO)(L-L)(PR3)2]ClO4 (L-L = bipy or phen) and [Rh(CO)(PR3)3]ClO4 with PR3 = P(p-YC6H4)3 (Y=Cl, F, Me or MeO) are described.  相似文献   

6.
Treatment of UO2X2 (X = OAc, Cl, NO3) with 1 mol equiv of (py)2CO in THF afforded the adducts [UO2X2{(py)2CO}] in almost quantitative yields. The same reactions in MeOH, in the presence of NEt3 for X = Cl and NO3, gave yellow crystals of [(UO2X)2{μ-(py)2C(OMe)O}2]·MeOH (X = OAc, 1·MeOH and X = Cl, 2·MeOH) and [{UO2(NO3)}2{μ-(py)2C(OMe)O}2] (3). Reactions of UO2X2 (X = OAc, Cl) with 2 mol equiv of (py)2CO and NEt3 in MeOH or further treatment of 1 and 2 with 1 mol equiv of (py)2CO and NEt3 afforded the methoxide derivative [{UO2(OMe)}2{μ-(py)2C(OMe)O}2] (4), while UO2(NO3)2 was transformed into [{UO2(NO3)}{UO2(OH)}{μ-(py)2C(OMe)O}2] (5). In these first structurally characterized actinide compounds with a (py)2CO-based ligand, the uranium atoms are located at the center of pentagonal (X = Cl and OMe) or hexagonal (X = OAc and NO3) bipyramids sharing one edge defined by the μ-alkoxo oxygen atoms. Crystals of [{UO2(OMe)}2{μ-(py)2C(OMe)O}2]·[(UO2)42-(py)2C(OMe)O}22-OAc)23-O)2(MeOH)2]·H2O (6·H2O) were serendipitously obtained in one experiment with [UO2(OAc)2(H2O)2] and (py)2CO.  相似文献   

7.
The use of methanol as solvent is essential for the formation of the double-bookshelf-type oxide cluster [(Cp*Rh)2Mo6O20(OMe)2]2− from [{Cp*Rh(μ-Cl)Cl}2] and four equivalents of [Mo2O7]2−. The reaction proceeds via [Cp*RhMo3O8(OMe)5]. The proposed structure for this key intermediate (shown schematically) is supported by electrospray ionization mass spectrometry and labeling experiments with CD3OD as solvent. Cp*=η5-C5Me5.  相似文献   

8.
The ready availability of rare parent amido d8 complexes of the type [{M(μ‐NH2)(cod)}2] (M=Rh ( 1 ), Ir ( 2 ); cod=1,5‐cyclooctadiene) through the direct use of gaseous ammonia has allowed the study of their reactivity. Both complexes 1 and 2 exchanged the di‐olefines by carbon monoxide to give the dinuclear tetracarbonyl derivatives [{M(μ‐NH2)(CO)2}2] (M=Rh or Ir). The diiridium(I) complex 2 reacted with chloroalkanes such as CH2Cl2 or CHCl3, giving the diiridium(II) products [(Cl)(cod)Ir(μ‐NH2)2Ir(cod)(R)] (R=CH2Cl or CHCl2) as a result of a two‐center oxidative addition and concomitant metal–metal bond formation. However, reaction with ClCH2CH2Cl afforded the symmetrical adduct [{Ir(μ‐NH2)(Cl)(cod)}2] upon release of ethylene. We found that the rhodium complex 1 exchanged the di‐olefines stepwise upon addition of selected phosphanes (PPh3, PMePh2, PMe2Ph) without splitting of the amido bridges, allowing the detection of mixed COD/phosphane dinuclear complexes [(cod)Rh(μ‐NH2)2Rh(PR3)2], and finally the isolation of the respective tetraphosphanes [{Rh(μ‐NH2)(PR3)2}2]. On the other hand, the iridium complex 2 reacted with PMe2Ph by splitting the amido bridges and leading to the very rare terminal amido complex [Ir(cod)(NH2)(PMePh2)2]. This compound was found to be very reactive towards traces of water, giving the more stable terminal hydroxo complex [Ir(cod)(OH)(PMePh2)2]. The heterocyclic carbene IPr (IPr=1,3‐bis(2,6‐diisopropylphenyl)imidazol‐2‐ylidene) also split the amido bridges in complexes 1 and 2 , allowing in the case of iridium to characterize in situ the terminal amido complex [Ir(cod)(IPr)(NH2)]. However, when rhodium was involved, the known hydroxo complex [Rh(cod)(IPr)(OH)] was isolated as final product. On the other hand, we tested complexes 1 and 2 as catalysts in the transfer hydrogenation of acetophenone with iPrOH without the use of any base or in the presence of Cs2CO3, finding that the iridium complex 2 is more active than the rhodium analogue 1 .  相似文献   

9.
The irreversible electroreduction of any of the three distinct dimolybdenum thiolate complexes [{MoCp(CO)}2(μ-SMe)3]+ and [{MoCp(CO)X}2(μ-SMe)2] (Cp = η5-C5H5; X = Cl or Br) gives a common intermediate which we identify as cis-[{MoCp(CO)}2(μ-SMe)2] (I). This intermediate slowly isomerises to the isolable product trans-[{MoCp(CO)}2(μ-SMe)2]; the isomerisation is catalysed by carbon monoxide, a process which we suggest takes place through CO binding across and elimination from the molybdenum-molybdenum double bond of (I).  相似文献   

10.
The synthesis and properties of neutral and cationic complexes of general formulae [{RhCl(diolefin)}2(CH2(pz)2)], [Rh(CO)2 (CH2(pz)2)][RhCl2(CO)2], (Rh(diolefin)(CH2(pz)2)]ClO2, [{Rh(diolefin)(PPh3)}2(CH2(pz)2)](ClO4)2, [Rh(CO)2(CH2(pz)2)]ClO4 and [Rh(CO)(CH2(pz)2)(PPh3)]ClO4 are described. The NMR spectra of [Rh(COD)(CH2(pz)2)]ClO4 complexes are discussed. X-ray structural analysis of [Rh(COD)(CH2(Pz)2)]ClO4 · 12C2H4Cl2 is presented; the final R factor is 0.061 for 2436 observed data, recorded with Cu-Kα, not corrected for absorption and with the sample inside a capillary. The Rh atom presents a distorted square planar coordination in a mononuclear arrangement. The COD ring has a twisted boat conformation, and the two halves of the CH2(Pz)2 moiety, which are quite similar to one another, form an angle of 47.2(4)°.  相似文献   

11.
The polyfunctional (H)PNX (X = O or N) ligands 1 and 2 react with [Rh(CO)2Cl]2 to give the corresponding chloro carbonyl complexes {Rh[κ2-(H)PN](CO)Cl} (1a and 2a), where the neutral ligands coordinate in a κ2-PN bidentate fashion, the square planar coordination being completed by the CO trans to N and the chloride trans to P. In chloroform solution 1a maintains its original structure, while 2a partially transforms into the cationic species {Rh[κ3-(H)PNO](CO)}Cl. The chloroform solutions of 1a and 2a react with AgPF6 to give the purely cationic species {Rh[κ3-(H)PNO](CO)}PF6 ([1a]+ and [2a]+), while addition of Et3N originates the neutral species {Rh[κ3-PNN′](CO)} (1b and 2b). All the complexes have been characterized by microanalysis, IR, 1H NMR as well as 31P{1H} NMR spectroscopy. The X-ray structures of ligand 1 and complex 1b are also reported.  相似文献   

12.
Summary Cleavage of [{Rh(diolefin)Cl}2] by bidentate heterocyclic chelating ligands (LL) has been studied, and diolefin neutral, ionic or ion-pair type compounds are formed depending on the ligands and/or the Rh: (LL) ratio employed. When the reactions are performed in media saturated with CO and with Rh: (LL)=21, only carbonylated ion-pair complexes are formed. The diolefin compounds react with tin(II) chloride yielding species containing trichlorostannato-groups. Subsequent reaction with CO leads to displacement of the diolefin and formation of the corresponding dicarbonyl species.  相似文献   

13.
The anions [Rh6(CO)15X]?, with X = COEt and CO(OMe), have been studied by single-crystal X-ray diffraction. They contain octahedral rhodium clusters, with mean metalmetal distances of 2.779 and 2.765 », respectively. The carbonyl stereochemistry in the two anions is similar to that of Rh6(CO)16, with one terminal CO group replaced by the X ligand. The RhC(carbomethoxy) bond distance (1.96(2) ») is significantly shorter than the RhC(acyl) distance (2.06(2) »).  相似文献   

14.
The reaction of Cp(CO)2FeEMe2 (E  As, Sb, Bi) with Me3P, Et3P, Me2PhP and (MeO)3P leads to a CO/R3P exchange and formation of the chiral derivatives Cp(CO)(R3P)FeEMe2. Cp(CO)[(MeO)3P]FeEMe2 rearranges already at room temperature to Cp(CO)[(Me3E]FeP(O)(OMe)2 which is transformed by (MeO)3P to Cp(CO)[(MeO)3P]FeP(O)(OMe)2. The high nucleophilicity of the new organometallic Lewis bases is established by the easy conversion of Cp(CO)(Me3P)FeSbMe2 to [Cp(CO)(Me3P)Fe(SbMe3)]I with MeI, or to [Cp(CO)(Me3P)FeSbMe2Fe(CO)LCp]Hal (L  CO, Hal  Cl; L  Me3P, Hal  Br) with Cp(CO)LFe-Hal, respectively. The new compounds are characterized by spectroscopy and elementary analyses.  相似文献   

15.
The kinetics and mechanisms of propadiene polymerization under the influence of [Rh(CO)2Cl]2, Rh(CO)2P(C6H5)3Cl, Rh(CO)3Cl are reported. The reaction rates are first-order in Rh(CO)2P(C6H5)3Cl and Rh(CO)3Cl and half-order in [Rh(CO)2Cl]2. They are second-order in the substrate for Rh(CO)3Cl and [Rh(CO)2Cl]2 and first-order for Rh(CO)2P(C6H5)3Cl. The data are interpreted in terms of a common intermediate mechanism. The formation of this common intermediate is the rate-determining step. A solvent effect is also discussed.  相似文献   

16.
The use of a strategy combining ligand design and changes of reaction conditions has been investigated with the goal of directing the assembly of mononuclear, dinuclear, tetranuclear, and polymeric copper(II) complexes. As a result, closely related copper monomers, alkoxo dimers, and hydroxo cubanes, along with a carbonate-bridged polymeric species, have been synthesized using the rigid, aliphatic amino ligands cis-3,5-diamino-trans-hydroxycyclohexane (DAHC), cis-3,5-diamino-trans-methoxycyclohexane (DAMC), and the glutaryl-linked derivative glutaric acid bis-(cis-3,5-diaminocyclohexyl) ester (GADACE). The composition of the monomeric complex has been determined by X-ray crystallography as [Cu(DAHC)2](ClO4)2 (1), the two dimers as [{Cu(DAHC)(OMe)}2](ClO4)2.MeOH (2) and [{Cu(DAMC)(OMe)(ClO4)}2] (3), the three Cu4O4 cubanes as [{Cu(DAHC)(OH)}4](ClO4)(4).2.5MeOH (4), [{Cu(DAMC)(OH)}4](ClO4)4.H2O (5), and [{Cu2(OH)2(GADACE)}2]Cl4.2MeOH.6H2O (6), and an infinite-chain structure as [{Cu(DAHC)(CO3)}n] (7). Furthermore, the cubane structures 4 and 5 have been investigated magnetically. Our studies indicate that formation of the monomeric, dimeric, and tetranuclear DAHC and DAMC complexes can be controlled by small changes in reaction conditions and that further preorganization of the ligand moiety by linking the DAHC cores (GADACE) allows more effective direction of the self-assembly of the Cu4O4 cubane core.  相似文献   

17.
The reaction of CpMo(CO)3X (X = Cl, I) with SbF5 in toluene leads to the cationic, halogen bridged compounds [{CpMo(CO)3}2X]SbF6 ( 1 , 2 ). CpW(CO)3Cl reacts with SbF5 to yield [{CpW(CO)3}2Cl]SbF6 ( 3 ), whereas with SbCl5, the oxidative substitution product [{CpW(CO)2Cl2}2Cl]SbCl6 ( 4 ) is formed, which decomposes in solution to yield the trichloride CpW(CO)2Cl3 ( 5 ). The strong Lewis acid SbF5 separates the halide from CpM(CO)3X (M = Mo, W), and the resulting cationic fragment “CpM(CO)3+” reacts with a further CpM(CO)3X forming a halonium bridge ( 1 – 3 ). The exclusive formation of SbF6 can be explained by the oxidizing power of SbF5. The IR, MS and NMR spectra of the compounds 1 – 5 as well as the X‐ray structure analysis of 5 are reported and discussed.  相似文献   

18.
Hemichelation is emerging as a new mode of coordination where non‐covalent interactions crucially contribute to the cohesion of electron‐unsaturated organometallic complexes. This study discloses an unprecedented demonstration of this concept to a Group 9 metal, that is, RhI. The syntheses of new 14‐electron RhI complexes were achieved by choosing the anti‐[(η66‐fluorenyl){Cr(CO)3}2] anion as the ambiphilic hemichelating ligand, which was treated with [{Rh(nbd)Cl}2] (nbd=norbornadiene) and [{Rh(CO)2Cl}2]. The new T‐shaped RhI hemichelates were characterized by analytical and structural methods. Investigations using the methods of the DFT and electron‐density topology analysis (NCI region analysis, QTAIM theory) confirmed the closed‐shell, non‐covalent and attractive characters of the interaction between the RhI center and the proximal Cr(CO)3 moiety. This study shows that, by appropriate tuning of the electronic properties of the ambiphilic ligand, truly coordination‐unsaturated RhI complexes can be synthesized in a manageable form.  相似文献   

19.
Solvothermal reaction of [MnCl2(terpy)] with elemental As and Se at a 1:1:2 molar ratio in H2O/trien (10:1) at 150 °C affords the linear trimanganese(II) complex [{Mn(terpy)}3(μ‐AsSe4)2] ( 1 ). The tridentate [AsSe2(Se2)]3? anions of 1 chelate the terminal {Mn(terpy)}2+ fragments and bridge these through their remaining Se atom to the central {Mn(terpy)}2+ moiety. Weak interactions of Mn1···Se and Mn3···Se bonds with length 2.914(7) and 3.000(7) Å link the molecules of 1 into infinite chains. Treatment of [MnCl2(cyclam)]Cl with As and Se at a 1:1:2 molar ratio in superheated H2O/CH3OH (1:1) at 150 °C yields the dinuclear complex [{Mn(cyclam)}2 (μ‐As2Se6)] ( 2 ), whose novel [(AsSe2)2(μ‐Se2)]4? ligands bridge the MnII atoms in a μ‐1κ2Se1, Se2: 2κ2Se5,Se6 manner.  相似文献   

20.
The complex mer-trans-[Mn(CO)3{P(OMe)2Ph}2X] (X = Cl, Br) is an intermediate in the conversion of fac-[Mn(CO)3{P(OMe)2,Ph}2,X] into mer- cis-[Mn(CO)2{P(OMe)2Ph}3X] in the presence of P(OMe)2Ph in benzene. No direct route between the latter two complexes could be detected kinetically. The results imply a trans carbonyl disposition as a prerequisite for higher carbonyl substitution in octahedral Mn1 carbonyl complexes.  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