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 31 毫秒
1.
The impetus for this work was the structure of a trinuclear complex with two carbonyl groups showing incipient triple bridging - Cp2Rh3(CO)4?. Its structure, barrier to rotation of one Rh(CO)2? piece vs. the rest of the molecule, and the nature of the bridging carbonyl interaction are analyzed. Isolobal analogies form an interesting connection between this complex and a bridged isomer of the recently synthesized carbene complexes, Cp2Rh2(CO)2CR2, one isomer of Cp2Rh3(CO)3, and hypothetical carbyne complexes Cp2Rh2(CO)2CH+,?. A general bonding model for Cp2Rh2(μ-CO)2X complexes is constructed. The model, rich in geometrical detail, allows minima for the bridging carbonyl groups bending toward and away from the bonded ligand X.  相似文献   

2.
Multinuclear NMR data (13C, 31P, 13C–{31P}, 13C–{103Rh} and 31P–{103Rh}) for a series of mono- and di-substituted derivatives of Rh6(CO)16 containing neutral two electron donor ligands [Rh6(CO)15L, (L=NCMe, py, cyclooctene, PPh3, P(OPh)3,1/2(μ2,η1:η1-dppe)); Rh6(CO)14(LL), (LL=cis-CH2=CMe-CMe=CH2, dppm, dppe, (P(OPh)3)2)] are reported; these data show that the solid state structure is maintained in solution. Detailed assignments of the 13CO NMR spectra of Rh6(CO)15(PPh3) and Rh6(CO)14(dppm) clusters have been made on the basis 13C–{103Rh} double resonance measurements and the specific stereochemical features of the observed long range couplings in these clusters have been studied. The stereochemical dependence of 3J(P–C) for terminal carbonyl ligands is discussed and the values of 3J(P–C) are found to be mainly dependent on the bond angles in the P–Rh–Rh–C fragment; these data enable the fine structure of the complex multiplets in the 13C–{1H} and 31P–{1H} NMR spectra of Rh6(CO)14 (dppm) to be simulated. Variable temperature 13C–{1H} NMR measurements on Rh6(CO)15(PPh3) reveal the carbonyl ligands in this complex to be fluxional. The fluxional process involves exchange of all the CO ligands except the two terminal CO's associated with the rhodium trans to the substituted rhodium and can be explained by a simple oscillation of the PPh3 on the substituted rhodium atom aided by concomitant exchange of the unique terminal CO on this rhodium with adjacent μ3-CO's.  相似文献   

3.
The Cp2Rh2(CO)n (n = 4, 3, 2, 1) derivatives have been examined by density functional theory using the BP86 and MPW1PW91 functionals. The known tricarbonyl Cp2Rh2(CO)3 is predicted to have a singly bridged structure with a predicted Rh–Rh single bond distance of ~2.70 Å in close agreement with the experimental value of 2.68 Å, determined by X-ray crystallography. In contrast to the cobalt analog, no evidence for a triply bridged Cp2Rh2(μ-CO)3 structure was found. The known dicarbonyl Cp2Rh2(CO)2 is predicted to have a doubly bridged structure with a predicted RhRh double bond distance of 2.58 Å in close agreement with the experimental RhRh double bond distance of 2.564 Å, found by X-ray crystallography for the permethylated derivative (η5-Me5C5)2Rh2(μ-CO)2. The monocarbonyl Cp2Rh2(CO) is predicted to have a four-electron donor bridging carbonyl group with a Rh–O distance of ~2.5 Å and a RhRh double bond distance of ~2.54 Å. This differs from Cp2Co2(CO) which was previously predicted to have only a two-electron donor bridging carbonyl group with a long Co?O distance and a short CoCo distance of ~2.0 Å suggesting a formal triple bond. For Cp2Rh2(CO)4 doubly bridged trans and cis isomers were found within ~1.0 kcal/mol in energy with non-bonding Rh?Rh distances of ~3.2 Å. However, these Cp2Rh2(CO)4 isomers are predicted to be unstable with respect both to CO loss to give Cp2Rh2(CO)3 and to fragmentation into two CpRh(CO)2 units.  相似文献   

4.
The 86-electron dicationic octahedral rhodium cluster [Rh6Cp6(μ6-C)]2+(PF6 -)2 containing Cp ligands and the interstitial carbon atom was synthesized by the reaction of Rh3Cp3(μ-CO)3 with RhCp(C2H4)2 Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp. 395–396, February, 1999  相似文献   

5.
Summary The following mono-, di- and tetra-nuclear rhodium carbonyl derivatives with silanolato and disilanolato ligands were prepared: Rh(CO)(PR3)2(OSiR3), [Rh(CO)2(OSiR3)]2, Rh2(CO)3 (PBu3)(OSiR3)2 and Rh4(CO)8(OSiPh2)2 (R = Me or Ph; R = n-Bu or Ph). The complexes may serve as models for the bonding of transition metals on silica surfaces.  相似文献   

6.
The reaction of rhodium(I) carbonyl chloride, [Rh(CO)2Cl]2, with dichromate, cerium(IV) sulfate, hexachloroplatinic acid or p-benzoquinone in aqueous hydrochloric acid proceeds by consumption of 4 equivalents of oxidizing agent per mole or rhodium(I) in accordance with the equation RhI(CO)2  4e + H2O → RhIII(CO) + 2H+ + CO2A “cyclic” oxidation mechanism is suggested.  相似文献   

7.
The anionic rhodium carbonyl clusters [Rh7(CO)16]3− and [Rh14(CO)25]4− can be easily prepared by a new simple and high yield one-pot synthesis starting from RhCl3·nH2O dissolved in ethylene glycol and involving two steps: (i) treatment of RhCl3·nH2O under 1 atm of CO at 50 °C to give [Rh(CO)2Cl2]; (ii) addition of a base (CH3CO2Na or Na2CO3) followed by reductive carbonylation under 1 atm of CO at an adequate temperature (50 °C for [Rh7(CO)16]3−; 150 °C for [Rh14(CO)25]4−). These new syntheses are more convenient than those previously reported, especially since such clusters are not accessible via silica surface-mediated reactions. This different behavior is due to the particular stabilization on the silica surface and under 1 atm of CO of an anionic carbonyl cluster, called A, which does not allow the formation of a higher nuclearity carbonyl cluster, called B, which was shown to be the key-intermediate in the synthesis of [Rh14(CO)25]4− working in ethylene glycol solution. Although it was not possible to isolate crystals of A and B suitable for X-ray structural determination, a combination of cyclovoltammetry, one of the few examples so far available of the use of this technique for anionic rhodium carbonyl clusters, infrared spectroscopy and elemental analyses suggest that A and B are probably the never reported [Rh7(CO)14] and [Rh15(CO)28]3− clusters, respectively. In particular the tentative formulation of the two clusters was carried out by a non-conventional method based on the existence of a linear correlation between carbonyl frequencies of the main band and the [(charge/Rh atoms)/CO number] ratio.  相似文献   

8.
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   

9.
Electronic Structure of Rh4(CO)12, a Model for Linear- and Bridge-bonded CO on Rhodium Catalysts The electronic structure of the Rh4(CO)12 cluster containing both linear- and bridgebonded CO groups has been studied by the EHMO method and compared with that of the Ir4(CO)12 cluster with only linear-bonded CO ligands. The charge distribution shows a distinctly higher π-back donation for the bridge-bonded CO groups. This result is compared with experimental data such as bond lengths, force constants of the C? O stretching frequencies and XPS data. It allows further an interpretation of results of CO hydrogenation on supported rhodium catalysts and on rhodium model complexes.  相似文献   

10.
The reactions of hydrosilane and/or alkyne as well as isonitriles with rhodium and rhodium cobalt mixed metal carbonyl clusters, e.g., Rh4(CO)12 and Co2Rh2(CO)12, are studied. Novel mixed metal complexes, e.g., CoRh(CO)5 (HCCBu n ), (R3Si)2Rh(CO) n Co(CO)4, Rh(R–NC)4Co(CO)4, Co2Rh2(CO)10(HCCR), and Co2Rh2(CO)9(HCCBu n ), are synthesized and identified. The catalytic activities of these rhodium and rhodium-cobalt mixed metal complexes are examined in hydrosilyation, silylformylation, and novel silylcarbocyclization reactions. Possible mechanisms for these reactions are proposed and discussed.  相似文献   

11.
Oxidation of rhodium(I) carbonyl chloride, [Rh(CO)2Cl]2, with copper(II) acetate or isobutyrate in methanol solutions yields binuclear double carboxylato bridged rhodium(II) complexes with RhRh bonds, [Rh(μ-OOCRκO)(COOMeκC)(CO)(MeOH)]2, where R=CH3 or i-C3H7. According to X-ray data, surrounding of each rhodium atom in these complexes is close to octahedral and consists of another rhodium atom, two oxygens of carboxylato ligands, terminal carbonyl group, C-bonded methoxycarbonyl ligand, and axial CH3OH. Methoxycarbonyl ligand is shown to originate from CO group of the parent [Rh(CO)2Cl]2 and OCH3 group of solvent. N- and P-donor ligands L (p-CH3C6H4NH2, P(OPh)3, PPh3, PCy3) readily replace the axial MeOH yielding [Rh(μ-OOCRκO)(COOMeκC)(CO)(L)]2. The X-ray data for the complex with R=i-C3H7, L=PPh3 showed the same molecular outline as with L=MeOH. Electronic effects of axial ligands L on the spectral parameters of terminal carbonyl group are essentially the same as in the known series of rhodium(I) complexes (an increase of δ13C and a decrease of ν(CO) with strengthening of σ-donor and weakening of π-acceptor ability of L).  相似文献   

12.
It has been established that reductive complexation of functionalized benzofulvenes, which are readily prepared from commercially available indene and 2‐methylindene, with RhCl3 in ethanol affords the corresponding indenyl–rhodium(III) dichlorides bearing substituents at the 1‐ (H or CO2Et), 2‐ (H or Me), and 3‐ [CH2Ph or CH2(2‐MeOC6H4)] positions. The indenyl–rhodium(III) complexes bearing one ethoxycarbonyl group showed higher thermal stability and regioselectivity than our previously reported CpERhIII complex toward the oxidative [3+2] annulation of acetanilides with internal alkynes.  相似文献   

13.
Chemisorption of Rh4(CO)12 on to a highly divided silica (Aerosil “0” from Degussa), Leads to the transformation: 3 Rh4(CO)12 → 2 Rh6(CO)16 + 4 CO. Such an easy rearrangement of the cluster cage implies mobility of zerovalent rhodium carbonyl fragments on the surface. Carbon monoxide is a very efficient inhibitor of this reaction, and Rh4(CO)12 is stable as such on silica under a CO atmosphere. Both Rh4(CO)12 and Rh6(CO)16 are easily decomposed to small metal particles of higher nuclearity under a water atmosphere and to rhodium(I) dicarbonyl species under oxygen. From the RhI(CO)2 species it is possible to regenate first Rh4(CO)12 and then Rh6(CO)16 by treatment with CO (Pco ? 200 mm Hg) and H2O (PH2O ? 18 mm Hg). The reduction of RhI(CO)2 surface species by water requires a nucleophilic attack to produce an hypothetical [Rh(CO)n]m species which can polymerize to small Rh4 or Rh6 clusters in the presence of CO but which in the absence of CO lead to metal particles of higher nuclearity. Similar results are obtained on alumina.  相似文献   

14.
The reaction of Rh4(CO)12 with Pd(PBu t 3)2 yielded the high nuclearity bimetallic hexarhodium-tripalladium cluster complex Rh6(CO)16[Pd(PBu t 3)]3, 10, in 11% yield. Compound 10 was converted to the hexarhodium-tetrapalladium cluster Rh6(CO)16[Pd(PBu t 3)]4, 11, in 62% yield by reaction with an additional quantity of Pd(PBu t 3)2. Both compounds were characterized crystallographically. Structurally, both compounds consist of an octahedral cluster of six rhodium atoms with sixteen carbonyl ligands analogous to that of the known compound Rh6(CO)16. Compound 10 also contains three Pd(PBu t 3) groups that bridge three Rh–Rh bonds along edges of the Rh6 octahedron to give an overall D3 symmetry to the Rh6Pd3 cluster. Compound 11 contains four edge bridging Pd(PBu t 3) groups distributed across the Rh6 octahedron to give an overall D2d symmetry to the Rh6Pd4 cluster. Each Rh–Pd connection in both compounds contains a bridging carbonyl ligand that helps to stabilize the bond between the Pd(PBu t 3) groups and the Rh atoms. Both compounds can be regarded as Pd(PBu t 3) adducts of Rh6(CO)16.  相似文献   

15.
A number of earlier unknown tri- and tetranuclear organometallic clusters of group VIII transition metals was synthesized by the addition of coordinatively unsaturated species to a single metal-metal bond. A number of novel heteronuclear clusters, CpCp 2 Rhm2(μ−CO)33−CO), (Cp′=Cp, Cp*; M2=Ru2, Fe2; RuFe); Cp2Cp 2 * Rh2M23−CO)3 (M = Ru, Fe); , Cp3Cp*Rh3M(μ3−CO)23−d) (M = Ru, Fe); Cp2Cp 2 *> Rh2Co23−CO)2,etc., was obtained. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 3, pp. 579–586, March, 1997.  相似文献   

16.
The structures of palladium carbonylcarboxylate clusters [Pd2(CO)2(RCOO)2] n (n = 2, R = CH3, CH2Cl, CF3, n = 3, R = CMe3, CHMe2, n-C5H11) are studied in benzene and tetrahydrofuran solutions by IR and 1H and 13C NMR spectroscopy. The clusters in the solid state have a planar cyclic metal framework with pairs of the carbonyl and carboxylate ligands alternately coordinated on its sides. In solutions, compounds under consideration contain one-type carbonyl ligands and one-type carboxylate ligands; their structures are similar to thaso in the solid state.  相似文献   

17.
2-(2-Trimethoxysilylethyl)pyridine, together with 3-methcryloxypropyltrimethoxysilane, was used to prepare a series of rhodium carbonyl complexes bound to silica via a pyridine group. The rhodium complex Rh2(CO)4Cl2 (Rh2) was used as the starting compound, and the immobilized complexes were prepared by four different routes which yielded both surface-bonded complexes and complexes bonded within the silicate matrix. These complexes were efficient catalysts of hydrosilylation of octene by triethxysilane. All the immobilized complexes were more than their homogeneous analogues and some could be re-used.  相似文献   

18.
The magnetic properties of molecular metal cluster compounds resemble those of small metal particles in the metametallic size regime. Even-electron metal carbonyl clusters with 10 or more metal atoms are paramagnetic, because their frontier orbital separations of less than 1 eV lead to high-spin electronic configurations. The rhodium cluster [Rh17S2(CO)32]3? gives EPR below 200 K withg=2.04, the first example of this type of paramagnetism in an even-electron carbonyl cluster of this 4d metal. Its spectral parameters are compared with those of osmium carbonyl clusters and some significant differences highlighted. Attempts have also been made to generate radical cations from lower-nuclearity, diamagnetic molecular clusters such as Rh6(CO)16 by chemical oxidation in sulphuric acid. An EPR active species (g=2.09) believed to be [Rh6(CO)16]+ has been obtained.  相似文献   

19.
A complete NMR study involving both 1D and 2D 13C-{103Rh} and 31P-{103Rh} HMQC measurements, on [Rh6C(CO)14(PPh3)]2- are reported and discussed, together with the multiple Rh quantum effects found for resonances associated with edge- and face-bridging CO's. As found in [Rh6C(CO)15]2-, the carbonyl ligands in [Rh6C(CO)14(PPh3)]2- undergo CO-intermolecular exchange with 13CO at different rates; for the edge-bridging CO's, the lower the value of 1J(Rh–CO), the faster the rate of intermolecular exchange with 13CO.  相似文献   

20.
Ferrocene‐amide‐functionalized 1,8‐naphthyridine (NP) based ligands {[(5,7‐dimethyl‐1,8‐naphthyridin‐2‐yl)amino]carbonyl}ferrocene (L1H) and {[(3‐phenyl‐1,8‐naphthyridin‐2‐yl)amino]carbonyl}ferrocene (L2H) have been synthesized. Room‐temperature treatment of both the ligands with Rh2(CH3COO)4 produced [Rh2(CH3COO)3(L1)] ( 1 ) and [Rh2(CH3COO)3(L2)] ( 2 ) as neutral complexes in which the ligands were deprotonated and bound in a tridentate fashion. The steric effect of the ortho‐methyl group in L1H and the inertness of the bridging carboxylate groups prevented the incorporation of the second ligand on the {RhII–RhII} unit. The use of the more labile Rh2(CF3COO)4 salt with L1H produced a cis bis‐adduct [Rh2(CF3COO)4(L1H)2] ( 3 ), whereas L2H resulted in a trans bis‐adduct [Rh2(CF3COO)3(L2)(L2H)] ( 4 ). Ligand L1H exhibits chelate binding in 3 and L2H forms a bridge‐chelate mode in 4 . Hydrogen‐bonding interactions between the amide hydrogen and carboxylate oxygen atoms play an important role in the formation of these complexes. In the absence of this hydrogen‐bonding interaction, both ligands bind axially as evident from the X‐ray structure of [Rh2(CH3COO)2(CH3CN)4(L2H)2](BF4)2 ( 6 ). However, the axial ligands reorganize at reflux into a bridge‐chelate coordination mode and produce [Rh2(CH3COO)2(CH3CN)2(L1H)](BF4)2 ( 5 ) and [Rh2(CH3COO)2(L2H)2](BF4)2 ( 7 ). Judicious selection of the dirhodium(II) precursors, choice of ligand, and adaptation of the correct reaction conditions affords 7 , which features hemilabile amide side arms that occupy sites trans to the Rh–Rh bond. Consequently, this compound exhibits higher catalytic activity for carbene insertion to the C?H bond of substituted indoles by using appropriate diazo compounds, whereas other compounds are far less reactive. Thus, this work demonstrates the utility of steric crowding, hemilability, and hydrogen‐bonding functionalities to govern the structure and catalytic efficacyof dirhodium(II,II) compounds.  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