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 468 毫秒
1.
The isolation of simultaneously low-coordinate and low-valent compounds is a timeless challenge for preparative chemists. This work showcases the preparation and full characterization of tri-coordinate rhodium(-I) and rhodium(0) complexes as well as a rare rhodium(I) complex. Reduction of [{Rh(μ-Cl)(IPr)(dvtms)}2] ( 1 , IPr=1,3-bis(2,6-diisopropylphenyl)imidazolyl-2-ylidene; dvtms=divinyltetramethyldisiloxane) with KC8 gave the trigonal complexes K[Rh(IPr)(dvtms)] and [Rh(IPr)(dvtms)], whereas the cation [Rh(IPr)(dvtms)]+ results from their oxidation or by abstraction of chloride from 1 with silver salts. The paramagnetic Rh0 complex is a unique fully metal-centered radical with the unpaired electron in the dz2 orbital. The Rh(-I) complex reacts with PPh3 with replacement of the NHC ligand, and behaves as a nucleophile, which upon reaction with [AuCl(PPh3)] generates the trigonal pyramidal complex [(IPr)(dvtms)Rh-Au(PPh3)] with a metal–metal bond between two d10 metal centers.  相似文献   

2.
A series of novel quasi‐scorpionate CNC donor ligands, MeC(2‐C5H4N){CH2(imidazole‐R)} (R = methyl, n‐butyl, n‐propenyl), in which a chelating bis(NHC) core is supplemented by a hemi‐labile pyridyl donor, were prepared. The coordination chemistry of these ligands was investigated with silver, palladium, rhodium and iridium. The single crystal X‐ray structures of [Rh(NC2Me)(COD)]Cl 8a and [Ir(NC2Pr)(COD)]Br 9b were determined. The catalytic potential of the rhodium and iridium complexes was assessed in the transfer hydrogenation of ketones; the iridium complexes, which show superior performance, form very effective and stable catalysts. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

3.
Palladium-catalysed Buchwald–Hartwig amination of ortho-substituted hindered aryl bromides or chlorides with 9H-carbazole has been investigated. In the amination of 1-bromo- or chloronaphthalene with 9H-carbazole, the combined use of Pd2(dba)3 as a Pd precursor, Buchwald ligands with two tert-butyl groups and LiOtBu or lithium hexamethyldisilazide as a base led to satisfactory yields. N,N’-Bis[2,6-bis(diphenylmethyl)-4-methoxyphenyl]imidazol-2-ylidene (IPr*OMe), which is a bulky N-heterocyclic carbene ligand, showed similar activity as Buchwald ligands with two tert-butyl groups. In contrast, only IPr*OMe provided satisfactory yields in the amination of 2-bromo-1,1′-biphenyl with 9H-carbazole. The amination of 2-bromo- or chlorotoluene and 1-(2-bromo- or chlorophenyl)naphthalene with 9H-carbazole proceeded smoothly when the IPr*OMe ligand was used.  相似文献   

4.
The 1,5-bis(3,5-dimethyl-1-pyrazolyl)-3-thiapentane ligand (bdtp) reacts with [Rh(COD)(THF)2][BF4] to give [Rh(COD)(bdtp)][BF4] ([1][BF4]), which is fluxional in solution on the NMR time scale. Its further treatment with carbon monoxide leads to a displacement of the 1,5-cyclooctadiene ligand, generating a mixture of two complexes, namely, [Rh(CO)2(bdtp)][BF4] ([2][BF4]) and [Rh(CO)(bdtp3N,N,S)][BF4] ([3][BF4]). In solution, [2][BF4] exists as a mixture of two isomers, [Rh(CO)2(bdtp2N,N)]+ ([2a]+) and [Rh(CO)2(bdtp3N,N,S)]+ ([2b]+; major isomer) rapidly interconverting on the NMR time scale. At room temperature, [2][BF4] easily loses one molecule of carbon monoxide to give [3][BF4]. The latter is prone to react with carbon monoxide to partially regenerate [2][BF4]. The ligands 1,2-bis[3-(3,5-dimethyl-1-pyrazolyl)-2-thiapropyl]benzene (bddf) and 1,8-bis(3,5-dimethyl-1-pyrazolyl)-3,6-dithiaoctane (bddo) are seen to react with two equivalents of [Rh(COD)(THF)2][BF4] to give the dinuclear complexes [Rh2(bddf)(COD)2][BF4]2 ([4][BF4]2) and [Rh2(bddo)(COD)2][BF4]2 ([5][BF4]2), respectively. In such complexes, the ligand acts as a double pincer holding two rhodium atoms through a chelation involving S and N donor atoms. Bubbling carbon monoxide into a solution of [4][BF4]2 results in loss of the COD ligand and carbonylation to give [Rh2(bddf)(CO)4][BF4]2 ([6][BF4]2). The single-crystal X-ray structures of [3][CF3SO3], [5][BF4]2 and [6][BF4]2 are reported.  相似文献   

5.
The first representative of chiral P-monodentate phosphite ligands containing quaternary ammonium substituents and its rhodium complex [Rh(COD)(L)2]BF4 (COD is 1,5-cycloctadiene) were obtained. The use of this ligand in Rh-catalyzed asymmetric hydrogenation of prochiral methyl esters of unsaturated acids allowed one to achieve optical yields up to 99%. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1395–1398, August, 2006.  相似文献   

6.
The coordination ability of the [(ppy)Au(IPr)]2+ fragment [ppy = 2-phenylpyridine, IPr = 1,3-bis(2,6-di-isopropylphenyl)-imidazol-2-ylidene] towards different anionic and neutral X ligands (X = Cl, BF4, OTf, H2O, 2-butyne, 3-hexyne) commonly involved in the crucial pre-equilibrium step of the alkyne hydration reaction is computationally investigated to shed light on unexpected experimental observations on its catalytic activity. Experiment reveals that BF4 and OTf have very similar coordination ability towards [(ppy)Au(IPr)]2+ and slightly less than water, whereas the alkyne complex could not be observed in solution at least at the NMR sensitivity. Due to the steric hindrance/dispersion interaction balance between X and IPr, the [(ppy)Au(IPr)]2+ fragment is computationally found to be much less selective than a model [(ppy)Au(NHC)]2+ (NHC = 1,3-dimethylimidazol-2-ylidene) fragment towards the different ligands, in particular OTf and BF4, in agreement with experiment. Effect of the ancillary ligand substitution demonstrates that the coordination ability of Au(III) is quantitatively strongly affected by the nature of the ligands (even more than the net charge of the complex) and that all the investigated gold fragments coordinate to alkynes more strongly than H2O. Remarkably, a stabilization of the water-coordinating species with respect to the alkyne-coordinating one can only be achieved within a microsolvation model, which reconciles theory with experiment. All the results reported here suggest that both the Au(III) fragment coordination ability and its proper computational modelling in the experimental conditions are fundamental issues for the design of efficient catalysts.  相似文献   

7.
New isocyanide ligands with meta‐terphenyl backbones were synthesized. 2,6‐Bis[3,5‐bis(trimethylsilyl)phenyl]‐4‐methylphenyl isocyanide exhibited the highest rate acceleration in rhodium‐catalyzed hydrosilylation among other isocyanide and phosphine ligands tested in this study. 1H NMR spectroscopic studies on the coordination behavior of the new ligands to [Rh(cod)2]BF4 indicated that 2,6‐bis[3,5‐bis(trimethylsilyl)phenyl]‐4‐methylphenyl isocyanide exclusively forms the biscoordinated rhodium–isocyanide complex, whereas less sterically demanding isocyanide ligands predominantly form tetracoordinated rhodium–isocyanide complexes. FTIR and 13C NMR spectroscopic studies on the hydrosilylation reaction mixture with the rhodium–isocyanide catalyst showed that the major catalytic species responsible for the hydrosilylation activity is the Rh complex coordinated with the isocyanide ligand. DFT calculations of model compounds revealed the higher affinity of isocyanides for rhodium relative to phosphines. The combined effect of high ligand affinity for the rhodium atom and the bulkiness of the ligand, which facilitates the formation of a catalytically active, monoisocyanide–rhodium species, is proposed to account for the catalytic efficiency of the rhodium–bulky isocyanide system in hydrosilylation.  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.
New chiral and non-chiral rhodium(I)–NHC complexes were synthesized. The first attempt by deprotonation of an imidazolinium salt with KOtBu and reaction with [Rh(COD)Cl]2 leads to the corresponding rhodium(I) complex. Due to the basic conditions during the reaction a loss of chirality occurs. An alternative transmetallation reaction with a silver(I)–NHC complex yields the desired rhodium(I)–NHC complex under retention of chirality. Both Rh complexes were fully characterized by analytical methods.  相似文献   

10.
Substitution of silver complex of 2‐chloro‐7‐(mesitylimidazolylidenylmethyl)naphthyridine (NpNHC) with palladium(II), rhodium(I) and iridium(I) metal precursors provided [Pd(C ,N ‐NpNHC)(η3‐allyl)](BF4) ( 5 ), RhCl(COD)(C ‐NpNHC) ( 6a ) and IrCl(COD)(C ‐NpNHC) ( 6b ), respectively. Abstraction of chloride from 6a and 6b with AgBF4 provided the chelation complexes [Rh(COD)(C ,N ‐NpNHC)](BF4) ( 7a ) and Ir(COD)(C ,N ‐NpNHC)(BF4) ( 7b ), respectively. All complexes were characterized using NMR and elemental analyses and the structural details of 5 and 6a were further confirmed using X‐ray crystallography. In catalytic activity studies, complex 5 was found to be an effective catalyst in the hydrogen‐transfer reduction of α,β‐unsaturated carbonyl compounds into the corresponding saturated carbonyl compounds.  相似文献   

11.
The reaction of zerovalent nickel compounds with white phosphorus (P4) is a barely explored route to binary nickel phosphide clusters. Here, we show that coordinatively and electronically unsaturated N‐heterocyclic carbene (NHC) nickel(0) complexes afford unusual cluster compounds with P1, P3, P5 and P8 units. Using [Ni(IMes)2] [IMes=1,3‐bis(2,4,6‐trimethylphenyl)imidazolin‐2‐ylidene], electron‐deficient Ni3P4 and Ni3P6 clusters have been isolated, which can be described as superhypercloso and hypercloso clusters according to the Wade–Mingos rules. Use of the bulkier NHC complexes [Ni(IPr)2] or [(IPr)Ni(η6‐toluene)] [IPr=1,3‐bis(2,6‐diisopropylphenyl)imidazolin‐2‐ylidene] affords a closo‐Ni3P8 cluster. Inverse‐sandwich complexes [(NHC)2Ni2P5] (NHC=IMes, IPr) with an aromatic cyclo‐P5? ligand were identified as additional products.  相似文献   

12.
Coordination chemistry of a new pyridine imidazole-2-ylidene ligand (pyNC) system with sterically hindered substituents toward rhodium(I) metal ions has been investigated. The rhodium complex [(pyNC)RhCl(COD)] (COD = 1,5-cyclooctadiene) was prepared via the transmetallation from the silver complex [(C-pyNC)2Ag]AgI2. Upon the abstraction of chloride, the pyridinyl nitrogen coordinated to the metal center and formed [(C,N-pyNC)Rh(COD)]BF4 with the chelation of pyNC. The pyridinyl nitrogen donor was found to be labile and could be replaced by various donors such as phosphine, azide and halides. Substitution of COD by various donors does not proceed except strong π-acid ligands such as CO and P(OCH3)3. However, the chelation of pyNC was replaced by the bisphosphine (P∼P) to form [(P∼P)2Rh]BF4, which was subsequently oxidized to yield [(P∼P)2Rh(O2)]BF4.  相似文献   

13.
A chiral N-heterocyclic carbene (NHC) ligand 1a bearing a m-terphenyl-based phosphate moiety as an anionic N-substituent has been developed. A rhodium complex [Rh(1a)(cod)]2 was synthesized and its structure was characterized by NMR and ESI-MS spectroscopy. This ligand gave high enantioselectivities in copper-catalyzed enantioselective boron conjugate additions to an α,β-unsaturated ester to give a chiral β-boryl ester.  相似文献   

14.
Multifaceted Coordination Chemistry of Vanadium(V): Substitution, Rearrangement Reactions, and Condensation Reactions of Oxovanadium(V) Complexes of the Tripodal Oxygen Ligand LOMe? = [η5‐(C5H5)Co{P(OMe)2(O)}3]? The octahedral oxovanadium(V) complex [V(O)F2LOMe] of the tripodal oxygen ligand LOMe? = [η5‐(C5H5)Co{P(OMe)2(O)}3]? reacts with alcohols and phenol with substitution of one fluoride ligand to form alkoxo complexes [V(O)F(OR)LOMe], R = Me, Et, i‐Prop, Ph. In the presence of water, however, both fluoride ions are substituted and a complex with the composition VO2LOMe can be isolated. The crystal structure shows that the oxo‐bridged trimer [{V(O)(LOMe)O}3] was synthesized. In the presence of BF3 the fluoride ligand in the alkoxo‐complex [V(O)F(OEt)LOMe] can be exchanged for pyridine to yield [V(O)(OEt)pyLOMe]BF4. Analogous attempts to exchange the fluoride ligand for tetrahydrofuran and acetonitrile induces a rearrangement reaction that leads to the vanadium complex [V(O)(LOMe)2]BF4. The crystal structure of this compound has been determined. Its 1H and 31P‐NMR spectra show that it is a highly fluxional vanadium complex at ambient temperature in solution. The two tripodal ligands LOMe? coordinate the vanadium centre as bidentate or tridentate ligands. The exchange bidentate/tridentate becomes slow on the NMR time scale below about 200 K.  相似文献   

15.
Stepwise introduction of the potential tripod ligands tris(3,5‐dimethyl‐1‐pyrazolyl)borate (Tp*) and tris(1‐cyclohepta‐2,4,6‐trienyl)phosphane into the coordination sphere of rhodium(I) leads mainly to [Tp*Rh{P(C7H7)3}] ( 4 ), in which Tp* is linked to the rhodium through a single pyrazolyl group and a non‐linear B–H–Rh bridge. This is the novel, now firmly established coordination mode κ2(N,B–H). The phosphane ligand is coordinated through one Rh–P and two Rh‐olefin bonds. Important structural features determined for the crystalline state of 4 are retained in solution, as shown by the 1H, 11B, 13C, 31P and 103Rh NMR spectra.  相似文献   

16.
The reaction of the rhodium(I) complexes [Rh(E)(PEt3)3] (E=GePh3 ( 1 ), H ( 6 ), F ( 7 )) with 1,1,3,3,3-pentafluoropropene afforded the defluorinative germylation products Z/E-2-(triphenylgermyl)-1,3,3,3-tetrafluoropropene and the fluorido complex [Rh(F)(CF3CHCF2)(PEt3)2] ( 2 ) together with the fluorophosphorane E-(CF3)CH=CF(PFEt3). For [Rh(Si(OEt)3)(PEt3)3] ( 4 ) the coordination of the fluoroolefin was found to give [Rh{Si(OEt)3}(CF3CHCF2)(PEt3)2] ( 5 ). Two equivalents of complex 2 reacted further by C−F bond oxidative addition to yield [Rh(CF=CHCF3)(PEt3)2(μ-F)3Rh(CF3CHCF2)(PEt3)] ( 9 ). The role of the fluorido ligand on the reactivity of complex 2 was assessed by comparison with the analogous chlorido complex. The use of complexes 1 , 4 and 6 as catalysts for the derivatization of 1,1,3,3,3-pentafluoropropene provided products, which were generated by hydrodefluorination, hydrometallation and germylation reactions.  相似文献   

17.
A general regioselective rhodium‐catalyzed head‐to‐tail dimerization of terminal alkynes is presented. The presence of a pyridine ligand (py) in a Rh–N‐heterocyclic‐carbene (NHC) catalytic system not only dramatically switches the chemoselectivity from alkyne cyclotrimerization to dimerization but also enhances the catalytic activity. Several intermediates have been detected in the catalytic process, including the π‐alkyne‐coordinated RhI species [RhCl(NHC)(η2‐HC?CCH2Ph)(py)] ( 3 ) and [RhCl(NHC){η2‐C(tBu)?C(E)CH?CHtBu}(py)] ( 4 ) and the RhIII–hydride–alkynyl species [RhClH{? C?CSi(Me)3}(IPr)(py)2] ( 5 ). Computational DFT studies reveal an operational mechanism consisting of sequential alkyne C? H oxidative addition, alkyne insertion, and reductive elimination. A 2,1‐hydrometalation of the alkyne is the more favorable pathway in accordance with a head‐to‐tail selectivity.  相似文献   

18.
The reaction of the trans‐hyponitrito complex [Ru2(CO)4(μ‐η2‐ONNO)(μ‐H)(μ‐PtBu2)(μ‐dppen)] ( 1 , dppen = Ph2PC(=CH2)PPh2) with tetrafluorido boric acid afforded the new complex salt [Ru2(CO)4(μ‐η2‐ONNOH)(μ‐H)(μ‐PtBu2)(μ‐dppen)]BF4 ( 2 ) containing the monoprotonate hyponitrous acid as the ligand in the cationic complex. Complex 1 showed a nucleophilic reactivity towards the trimethyloxonium cation resulting in the monoester derivative of the hyponitrous acid [Ru2(CO)4(μ‐η2‐ONNOMe)(μ‐H)(μ‐PtBu2)(μ‐dppen)]BF4 ( 3 ). During heating of compound 2 in ethanol under reflux for a short time nitrous oxide was liberated affording unexpectedly a new tridentate 2, 2‐bis(diphenylphosphanyl)ethanolato ligand formed by an intramolecular attack of an intermediate hydroxido ligand towards the unsaturated carbon carbon double bond in the bridging dppen ligand. Thus the complex salt [Ru2(CO)4{μ‐η3‐OCH2CH(PPh2)2}(μ‐H)(μ‐PtBu2)]BF4 ( 4 ) was formed in good yields. The new compounds 2 , 3 , and 4 were characterized by spectroscopic means as well as their molecular structures were determined in the crystal.  相似文献   

19.
The feasibility of oxidative addition of the P−H bond of PHPh2 to a series of rhodium complexes to give mononuclear hydrido-phosphanido complexes has been analyzed. Three main scenarios have been found depending on the nature of the L ligand added to [Rh(Tp)(C2H4)(PHPh2)] (Tp= hydridotris(pyrazolyl)borate): i) clean and quantitative reactions to terminal hydrido-phosphanido complexes [RhTp(H)(PPh2)(L)] (L=PMe3, PMe2Ph and PHPh2), ii) equilibria between RhI and RhIII species: [RhTp(H)(PPh2)(L)]⇄[RhTp(PHPh2)(L)] (L=PMePh2, PPh3) and iii) a simple ethylene replacement to give the rhodium(I) complexes [Rh(κ2-Tp)(L)(PHPh2)] (L=NHCs-type ligands). The position of the P−H oxidative addition–reductive elimination equilibrium is mainly determined by sterics influencing the entropy contribution of the reaction. When ethylene was used as a ligand, the unique rhodaphosphacyclobutane complex [Rh(Tp)(η1-Et)(κC,P-CH2CH2PPh2)] was obtained. DFT calculations revealed that the reaction proceeds through the rate limiting oxidative addition of the P−H bond, followed by a low-barrier sequence of reaction steps involving ethylene insertion into the Rh−H and Rh−P bonds. In addition, oxidative addition of the P−H bond in OPHPh2 to [Rh(Tp)(C2H4)(PHPh2)] gave the related hydride complex [RhTp(H)(PHPh2)(POPh2)], but ethyl complexes resulted from hydride insertion into the Rh−ethylene bond in the reaction with [Rh(Tp)(C2H4)2].  相似文献   

20.
Summary The rhodium(I) carbonyl compounds [Rh(CO)L22] [BF4]. 1/2CH2Clnn2 (L = PPh2 or AsPh3) react with the nucleophiles OMe, RCOO (R = Me, Et) under nitrogen to form [Rh(OR)(CO)L2] (1)–(2) and [Rh(OOCR)(CO)L2] (7)–(10), respectively. Addition of [Rh(CO)2(PPh3)2]-[BF 4] to OMe under nitrogen produces [Rh(COOMe)-(CO) (PPh3)2]-MeOH (3), whilst reactions of [Rh(CO)-(PPh3)2] [BF4]·1/2CH2Cl2 and [Rh(CO)2(PPh3)2] [BF4] with OR- (R = Me, Et or n-Pr) in the presence of CO produce [Rh(COOR)(CO)2(PPh3)2] (4)–(6). The products have been characterised by i.r., 1H, 31P, 13Cn.m.r. spectroscopy and elemental analysis.  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