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.
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   

2.
The crystal structures of Rh4(CO)10(PPh3)2 and Rh4(CO)9P(OPh)33 are reported. 31P-1H NMR studies on Rh4(CO)12-x {P(OPh)3}x(X  1, 2 and 3) show that each derivative exists as only one isomer in solution whereas the analogous triphenylphosphine derivatives can exist as different isomers. A quantitative redistribution of triphenylphosphites occurs on mixing Rh4-(CO)12-xLx with Rh4(CO)12-yLy (L  P(OPh)3; x  0, 1, 2, yx + 2; x  0, yx + 4) to give Rh4(CO)12-zLz[z12(x + y)]; a related rapid intermolecular randomisation of carbonyls occurs on mixing Rh4(12CO)12 with Rh4(13CO)12.  相似文献   

3.
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   

4.
The iridium and rhodium complexes [MCl(CO)2(NH2C6H4Me-4)] (M = Ir or Rh) react with [Os3(μ-H)2(CO)10] to give the tetranuclear clusters [MOs3(μ-H)2(μ-Cl)(CO)12]; the iridium compound being structurally identified by X-ray diffraction. Similarly, [IrCl(CO)2(NH2C6H4Me-4)] and [Rh2(μ-CO)2(η-C5Me5)2] afford the tetranuclear cluster [Ir2Rh2(μ-CO)(μ3-CO)2(CO)4(η-C5Me5)2], also characterised by single-crystal X-ray crystallog  相似文献   

5.
In the hydroformylation of ethylene with approximately equimolar H2/D2 mixtures and Rh4(CO)12 or Co2(CO)8 as the catalyst precursor about 50% of propionaldehyde-d1 was formed. The propionaldehyde-d0/d2 ratio was ~ 3 for rhodium and ~ 2.6 for the cobalt catalyst. On the basis of the results and assuming that there is no rapid M(H)2/M(D)2 scrambling, activation of hydrogen through M(H)2 or M(H)2(olefin) complexes can be excluded.  相似文献   

6.
The anion [Rh6(CO)14]4? has been isolated from the reaction of [Rh6(CO)16] with alkali hydroxides in aqueous solution. It shows high reactivity towards electrophiles and in redox condensations with other rhodium clusters; and is rapidly decomposed by carbon monoxide.  相似文献   

7.
The reaction of Rh(CO)2acac with triphenylantimony in the presence of cesium benzoate in tetraethylene glycol/dimethyl ether solution resulted in the selective formation of [Rh12Sb(CO)27]3- (66% yield) after 3 h of contact time under ≈400 atm of carbon monixide and hydrogen (CO/H2  1) at 140–160°C. The cluster has been isolated as the [Cs(18-Crown-6)2]+, [(CH3)4]+, [(C2H5)4N]+, (Ph3P)2N]+ and [PhCH2N(C2H5)3]+ salts. The [(C2H5)4N]3 [Rh12Sb(CO)27] complex has been characterized via a complete three-dimensional X-ray diffraction study. The complex crystallizes in the space group R3c with a  23.258(13) Å, c  22.811(4) Å, V  10 686 Å3 and p(calcd.)  2.334 g cm-3 for mol.wt. 2503.66 and Z  6. Diffraction data were collected with an Enraf-Nonius CAD 4 automated diffractometer using graphite-monochromatized Mo-Kα radiation. The structure was solved by direct methods and refined by difference-Fourier and least-squares techniques. All non-hydrogen atoms have been located and refined: final discrepancy indices are Rf  3.5% and Rwf  4.6% for 3011 reflections. The anion's structure consists of twelve rhodium atoms situated at the corners of a distorted icosahedron with contacts of 2.807(1), 2.861(1), 2.874(1), 2.999(1), 3.017(1) and 3.334(1) Å and rhodium—antimony contacts of 2.712(0) Å. Rhodium—rhodium bond distances of 2.807 and 3.017 Å are in the range usually found for these complexes although a distance of 3.334 Å may be longer than expected from bonding interactions. The sum of the covalent radii of antimony and rhodium, 2.80 Å, is intermediate between the two observed RhSb contacts. The anion cluster structure is that of distorted icosahedron. This polyhedron has previously been found in [B12H12]2- but not with transition metal clusters. A comparison between the structures of rhodium carbonyl clusters and boranes shows the occurrence of similar structural features. Applications of bonding theories based on the boranes, such as Wade's rules, to rhodium carbonyl clusters shows the extent in which these rules are obeyed.  相似文献   

8.
Reactions of Rh2(CO)4Cl2 with 1,5-cyclooctadiene (COD) and tetramethylallene (TMA) were performed separately in anhydrous hexane under argon atmosphere. Multiple perturbations of Rh2(CO)4Cl2, COD and TMA were also performed during the reactions. These two reactions were monitored by in-situ FTIR (FIR and MIR) and/or Raman spectroscopies and the collected spectra were further analyzed with BTEM family of algorithms. DFT calculations were performed to identify the organometallic species present. The known diene complex Rh2(CO)2Cl24-C8H12) and a new allene complex Rh2(CO)3Cl22-C7H12) were formed as the two primary organo-rhodium products. Their pure component spectra were reconstructed in the three characteristic regions of 200-680, 800-1360, and 1500-2200 cm−1. Their relative concentrations were also obtained by the least square fitting of the carbonyl region 1500-2200 cm−1. The present contribution shows the usefulness of combining in-situ spectroscopic measurements, BTEM analysis and DFT spectral prediction in order to analyze organometallic reactions at high dilution and identify the reaction products.  相似文献   

9.
We report a very efficient homogeneous system for the visible‐light‐driven hydrogen production in pure aqueous solution at room temperature. This comprises [RhIII(dmbpy)2Cl2]Cl ( 1 ) as catalyst, [Ru(bpy)3]Cl2 ( PS1 ) as photosensitizer, and ascorbate as sacrificial electron donor. Comparative studies in aqueous solutions also performed with other known rhodium catalysts, or with an iridium photosensitizer, show that 1) the PS1 / 1 /ascorbate/ascorbic acid system is by far the most active rhodium‐based homogeneous photocatalytic system for hydrogen production in a purely aqueous medium when compared to the previously reported rhodium catalysts, Na3[RhI(dpm)3Cl] and [RhIII(bpy)Cp*(H2O)]SO4 and 2) the system is less efficient when [IrIII(ppy)2(bpy)]Cl ( PS2 ) is used as photosensitizer. Because catalyst 1 is the most efficient rhodium‐based H2‐evolving catalyst in water, the performance limits of this complex were further investigated by varying the PS1 / 1 ratio at pH 4.0. Under optimal conditions, the system gives up to 1010 turnovers versus the catalyst with an initial turnover frequency as high as 857 TON h?1. Nanosecond transient absorption spectroscopy measurements show that the initial step of the photocatalytic H2‐evolution mechanism is a reductive quenching of the PS1 excited state by ascorbate, leading to the reduced form of PS1 , which is then able to reduce [RhIII(dmbpy)2Cl2]+ to [RhI(dmbpy)2]+. This reduced species can react with protons to yield the hydride [RhIII(H)(dmbpy)2(H2O)]2+, which is the key intermediate for the H2 production.  相似文献   

10.
Although very bulky ligands e.g.(o-MeC6H4)3E or (μ-C10H7)3E (E = P or As) are inert, the normal photochemical or thermal reaction of tertiary phosphines or arsines, L, with [Mn2(CO)10] is CO substitution with the formation of [Mn2(CO)8(L)2] derivatives (I). At elevated temperatures some triarylarsines, R3As, undergo Lambert's reaction with ligand fragmentation to give [Mn2(CO)8(μ-AsR2)2] complexes (II) (R = Ph, p-MeOC6H4, p-FC6H4, or p-CIC6H4) even though, in the absence of [Mn2(CO)10] R3As are stable under the same conditions. Exceptional behaviour is exhibited by (p-Me2NC6H4)3- As which forms a product of type I; by some HN(C6H4)2AsR which give a product of type II as a result of loss of the non-aryl groups R = PhCH2, cyclo-C6H11, or MeO; and by Ph(α-C10H72P which is the only phosphine to form a product of type II, albeit in trace amounts only. The thermal decomposition of a n-butanol solution of [Mn2(CO)8(AsPh3)2] in a sealed tube gives C6H6 and [Mn2(CO)8(α-AsPh2)2], whilst in an open system in the presence of various tertiary phosphines, L, [Mn(H)(CO)3(L)2] are obtained. It is suggested that Lambert's reaction is a thermal fragmentation of [Mn(CO)4(AsR3]* radicals, the first to be recognised. They lose the radical R* which abstracts hydrogen from the solvent. The resulting [Mn(CO)4(AsR2)] moiety dimerises to [Mn2(CO)8-(α-AsR2)2]. the reaction is facilitated by the stability of the departing radical (e.g. PhCH2 or MeO) and, as the crowding about As is relieved, by its size (e.g. Ph, cyclo-C6H11, o-MeC6H4, or α-C10H7). In general, phosphine-substituted radicals [Mn(CO)4(PR)3]* do not undergo this decomposition, probably because the PC bonds are much stronger than AsC.  相似文献   

11.
In this study selected bidentate (L2) and tridentate (L3) ligands were coordinated to the Re(I) or Tc(I) core [M(CO)2(NO)]2+ resulting in complexes of the general formula fac-[MX(L2)(CO)2(NO)] and fac-[M(L3)(CO)2(NO)] (M = Re or Tc; X = Br or Cl). The complexes were obtained directly from the reaction of [M(CO)2(NO)]2+ with the ligand or indirectly by first reacting the ligand with [M(CO)3]+ and subsequent nitrosylation with [NO][BF4] or [NO][HSO4]. Most of the reactions were performed with cold rhenium on a macroscopic level before the conditions were adapted to the n.c.a. level with technetium (99mTc). Chloride, bromide and nitrate were used as monodentate ligands, picolinic acid (PIC) as a bidentate ligand and histidine (HIS), iminodiacetic acid (IDA) and nitrilotriacetic acid (NTA) as tridentate ligands. We synthesised and describe the dinuclear complex [ReCl(μ-Cl)(CO)2(NO)]2 and the mononuclear complexes [NEt4][ReCl3(CO)2(NO)], [NEt4][ReBr3(CO)2(NO)], [ReBr(PIC)(CO)2(NO)], [NMe4][Re(NO3)3(CO)2(NO)], [Re(HIS)(CO)2(NO)][BF4], [99Tc(HIS)(CO)2(NO)][BF4], [99mTc(IDA)(CO)2 (NO)] and [99mTc(NTA)(CO)2(NO)]. The chemical and physical characteristics of the Re and Tc-dicarbonyl-nitrosyl complexes differ significantly from those of the corresponding tricarbonyl compounds.  相似文献   

12.
Solid solutions of the end members Fe2WO6, Cr2WO6, and Rh2WO6 have been prepared and their crystallographic and magnetic properties studied. All solid solutions crystallize with the trirutile structure, and their magnetic behavior is characterized by the existence of antiferromagnetic interactions and effective molar Curie constants corresponding to those expected from contributions of the spinonly moments of high-spin Fe3+, Cr3+, and diamagnetic low-spin Rh3+ ions. Fe2WO6 crystallizes with the tri-α-PbO2 structure and is antiferromagnetic and conducting. The random rutile Rh2WO6 is conducting, and the difference between its magnetic and electric properties and those of the inverse trirutile Cr2WO6 are discussed in terms of possible interactions between Cr3+(3d) or Rh3+(4d) orbitals and W6+(5d) orbitals.  相似文献   

13.
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   

14.
Rh4(CO)12 anchored on γ-Al2O3 (Rh4(CO)12/Al2O3) has been studied as a catalyst for the hydrogenation of 1,3-trans-pentadiene. Under mild conditions (1 atm H2 and temperatures between 60°C and 80°C) hydrogenation occurs at only one of the double bonds of the diene, and analysis of the products shows that the terminal double bond is preferentially hydrogenated. Hydrogenation of the second double bond of the conjugated diene occurring only after all the 1,3-trans-pentadiene has been consumed. In this respect Rh4(CO)12/Al2O3 behaves like toluene solutions of Rh4(CO)12. Anchoring of Rh4(CO)12 on the solid support gives a catalyst which is less active but more stable than toluene solutions of Rh4(CO)12. The effects of CO and of triphenylphosphine on catalytic activity and on specificity of Rh4(CO)12/Al2O3 have also been investigated and both shown to cause a reduction of the rate of hydrogenation of 1,3-trans-pentadiene.  相似文献   

15.
Acetone and methyl ethyl kektone undergo facile and direct metalation at the methyl groups by a cationic (octaethylporphyrinato) rhodium (III) complex with a non-coordinating perchlorate counteranion, (OEP)RhIII(ClO4), under mild conditions. Acetylacetone and ethyl acetoacetate are similarly metalated at the internal methylene groups. The metalation of acetone is firs-order with respect to both rhodium complex and ketone, and involves the (OEP)RhIII(ClO4)-assisted, rate-determining enolization of the ketone. The resulting 2-oxopropyl-rhodium derivative undergoes facile cleavage of the CRh bond with electrophiles such as H+ and Br2. When cyclohexanone is used as substrate, on the other hand, (OEP)RhIII(ClO4) catalyzes the aldol condensation of the ketone effectively, where the intermediate cyclohexanone enolate reacts with the ketone or other carbonyl compound present and regenerates the RhIII complex. An essential aspect of the present reaction is the remarkable ability of (OEP)RhIII(ClO4) to promote enolization of simple ketones by activation with charge-separated [(OEP)RhIII+ (a Lewis acid) under mild and neutral conditions. The second-order rate constant of (OEP)RhIII(ClO4)-assisted enolization of acetone at 30°C (k2 = 2.6 × 10−4 M−1 sec−1) is 107 times as large as that of its spontaneous enolization in water, where water is both acid and bse.  相似文献   

16.
The activation of the CN triple bond of benzonitrile in the presence of acetic acid and of Os3(CO)12 or H2Os3(CO)10 has been studied. When Os3(CO)12 reacts with PhCN and acetic acid in refluxing n-octane the three main products are (μ-H)Os3(CO)10(μ-O2CCH3) (I), (μ-H)Os3(CO)10(μ-NCHPh) (II) and (μ-H)Os3(CO)10(μ-NHCH2Ph) (III); II and III are analogues of (μ-H)Ru3(CO)10(μ-NCHPh) and (μ-H)Ru3(CO)10(μ-NHCH2Ph) obtained from PhCN, Ru3(CO)12 or H4Ru4(CO)]12, and acetic acid. In contrast to the reaction with ruthenium clusters, Os3(CO)12 and H2Os3(CO)10 also give the adduct Os3(CO)10(CH3COOH) (I). The structure of I has been fully elucidated by X-ray diffraction. Crystals of I are monoclinic, space group P21/m, with unit cell parameters a 7.858(6), b 12.542(8), c 9.867(6) Å, β 109.92(2)°, Z = 2. In I an edge of the triangular cluster of osmium atoms is doubly bridged by a hydride and an acetate ligand. Ten terminal carbonyl groups are bonded to the metal atoms.  相似文献   

17.
The formation of a new compound, the most characteristic IR absorption bands of which appear at 2007 cm-1 and 1956 cm-1, has been in the reaction between Co2(CO)8 and Rh4(CO)12 under carbon monoxide pressure in a hydrocarbon medium. The same compound is also formed either by the reaction of Co2(CO)8 with [Rh(CO)2Cl]2 or by the reaction of Co3Rh(CO)12 with carbon monoxide. The new complex has not been isolated in a pure state, but the formula CoRh(CO)7 is proposed on the basis of the stoichiometry of its formation and its physico-chemical properties. Equilibrium constants and thermo-dynamic parameters for the reaction 2 Co2(CO)8 + Rh4(CO)12  4 CoRh(CO)7 have been estimated. Possible structures for the new complex are discussed on the basis of its IR spectrum.  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.
Ethanolic solutions of RhIII chloride exposed to γ-radiation under CO atmosphere are shown to be totally reduced into RhI complexes (Rh2(CO)4Cl2 and Rh(CO)2Cl-2) within a few hours with a radiolytic reduction yield of about 6.0 elementary reductions/100 eV (6.2·10-7 mol·J-1). The chloride ions freed in the medium inhibit further reduction through Rh(CO)2Cl-2 formation. On addition of copper metal under the same conditions, RhIII is transformed into Rh6(CO)16 with a conversion yield 50%. This cluster is formed via Rh2(CO)4Cl2 although Rh(CO)2Cl-2 is also present under these conditions. Rh6(CO)16 cluster is also formed under radiolysis by direct reduction of Rh2(CO)4Cl2, but metallic rhodium and other reduced products are obtained at the same time.  相似文献   

20.
The preparation and characterization by elemental analysis, electronic and infrared spectroscopy are reported for the monomeric complexes cis-(amine)-M(CO)2Cl (M = Ir or Rh, amine = 1,8-naphthyridine or pyridazine; M = Ir, amine = o-phenylenediamine) and the binuclear species (1,8-naphthyridine)Rh2(CO)4Cl2, (1,8-naphthyridine)IrRh(CO)4Cl2, (pyrazine)Rh2(CO)4Cl2 and (1,3-di-4-pyridylpropane)Rh2(CO)4Cl2.  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