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.
Analogy with the isolable oxo cluster [Fe3(CO)93-O)]2−, which is structurally interesting and synthetically useful, prompted the present attempt to synthesize its ruthenium analog. Although the high reactivity of [Ru3(CO)93-O)]2− (I) prevented its isolation, the reaction of this species with [M(CO)3(NCCH3)]+, where M = Mn or Re, yields [PPN][MRu3(CO)1223-NC(μ-O)CH3]. The high nucleophilicity of the oxo ligand in [Ru3(CO)93-O)]2− (I) appears to be responsible for the conversion of acetonitrile to an acetamidediato ligand and for the instability of I. The crystal structure of [PPN][MnRu3(CO)1223-NC(μ-O)CH3)]] reveals a hinged butterfly array of metal atoms in which the acetamidediato ligand bridges the two wings with μ3-N bonding to an Mn and two Ru atoms, and μ-O bonding to an Ru atom.  相似文献   

2.
The dimetal μ-vinylidene complexes Cp(CO)2MnPt(μ-C = CHPh)L2 (L = tert.-phosphine or -phosphite), which have been obtained by coupling of the mononuclear complex Cp(CO)2Mn=C=CHPh and unsaturated PtL2 unit, add smoothly the Fe(CO)4 moiety to produce trimetal MnFePt compounds. The μ3-vinylidene cluster CpMnFePt(μ3-C=CHPh)(CO)6(PPh3) was prepared in quantitative yields from the reactions of Cp(CO)2MnPt(μ-C=CHPh)(PPh3)L (L = PPh3 or CO) with Fe2(CO)9 in benzene at 20 °C. The phosphite-substituted complexes Cp(CO)2Mnpt(μ-C=CHPh)L2 (L = P(OEt)3 or P(OPri)3) react under analogous conditions with Fe2(CO)9 to give mixtures (2:3) of the penta- and hexacarbonyl clusters, CpMnFePt(μ3-C = CHPh)(CO)5L2 and CpMnFePt(μ3-C = CHPh)(CO)6L, respectively. The similar reaction of the dimetal complex Cp(CO)2MnPt(μ-C = CHPh)(dppm), in which the Pt atom is chelated by dppm = Ph2PCH2PPhPin2 ligand, gives only a 15% yield of the analogous trimetal μ3-vinylidene hexacarbonyl product CpMnFePt(μ3-C = CHPh)(CO)(dppm), but the major product (40%) is the tetranuclear μ4-vinylidene cluster (dppm)PtFe34-C = CHPh)(CO)9. The IR and 1H, 13C and 31P NMR data for the new complexes are reported and discussed.  相似文献   

3.
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   

4.
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   

5.
Treatment of MHCl(CO)(PPh3)3 (M=Ru, Os) with (CH2=CH)SnR3 is a good general route to the coordinatively unsaturated osmium and ruthenium stannyl complexes M(SnR3)Cl(CO)(PPh3)2 (1: M=Ru, R=Me; 2: M=Ru, R = n-butyl; 3: M=Ru, R = p-tolyl; 4: M=Os, R=Me). These coordinatively unsaturated complexes readily add CO and CN-p-tolyl to form the coordinatively saturated compounds M(SnR3)Cl(CO)L(PPh3)2 (5: M=Ru, R=Me, L=CO; 6: M=;Ru, R = n-butyl, L=CO; 7: M=Ru, R = p-tolyl, L=CO; 8: M=Os, R=Me, L=CO; 9: M=Ru, R=Me, L=CN-p-tolyl; 10: M=Ru, R = n-butyl, L=CN-p-tolyl; 11: M=Os, R=Me, L=CN-p-tolyl). In addition, the chloride ligand in Ru(SnR3)Cl(CO)(PPh3)2 proves to be labile and treatment with the potentially bidentate anionic ligands, dimethyldithiocarbamate or diethyldithiocarbamate, affords the coordinatively saturated compounds Ru(SnR3)(η2-S2CNR′2)(CO)(PPh3)2 (12: R=Me, R′ = Me; 13: R=Me, R′ = Et; 14: R = n-butyl, R′ = Me; 15: R = p-tolyl, R′ = Me; 16: R = p-tolyl, R′ = Et). Chloride is also displaced by carboxylates forming the six-coordinate compounds Ru(SnR3)(η2-O2CR′)(CO)(PPh3)2 (17: R=Me, R′ = H; 18: R=Me, R′ = Me; 19: R=Me, R′ = Ph; 20: R = n-butyl, R′ = Me; 21: R = p-tolyl, R′ = Me). IR and 1H NMR spectral data for all the new compounds and 31P and 119Sn NMR spectral data for selected compounds are reported.  相似文献   

6.
Metathesis of [(η33−C10H16)Ru(Cl) (μ−Cl)]2 (1) with [R3P) (Cl)M(μ-Cl)]2 (M = Pd, Pt), [Me2NCH2C6H4Pd(μ-Cl)]2 and [(OC)2Rh(μ-Cl)]2 affords the heterobimetallic chloro bridged complexes (η33-C10H16) (Cl)Ru(μ-Cl)2M(PR3)(Cl) (M = Pd, Pt), (η33-C10H16) (Cl)Ru(μ-Cl)2PdC6H4CH2NMe2 and (η33-C10H16) (Cl)Ru(μ-Cl)2Rh(CO)2, respectively. Complex 1 reacts with [Cp*M(Cl) (μ-Cl)]2 (M = Rh, Ir), [p-cymene Ru(Cl) (μ-Cl]2 and [(Cy3P)Cu(μ-Cl)]2 to give an equilibrium of the heterobimetallic complexes and of educts. The structures of (η33-C10H16)Ru(μ-Cl)2Pd(PR3) (Cl) (R = Et, Bu) and of one diastereoisomer of (η33-C10H16)Ru(μ-Cl)2IrCp*(Cl) were determined by X-ray diffraction.  相似文献   

7.
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   

8.
Complexes Ru(CO)2 (CH=CHR) (C6H4X-4)L2 (R=tBu, Ph, OEt; X=H, Cl, OMe; L=PMe3, PMe2Ph, P(OMe)2Ph) in which the two phosphorus ligands are mutually cis (isomer 1) react readily with ligands tBuNC, CO and P(OMe)3 to give complexes in which one of the organic ligands has migrated onto a carbonyl ligand. Vinyl migration products (5) retain the mutually cis geometry of the phosphorus ligands, and are unstable: one of the decomposition products is the ketone RCH=CHC(O)C6H4X-4. Phenyl migration products (4) are stable and have the phosphorus ligands in mutually trans positions; an X-ray crystal structure of Ru(CO) (CNtBu) {C(O)Ph} (CH=CHPh) (PMe2Ph)2 was obtained. In both cases, the incoming ligand enters trans to the newly formed acyl ligand. Vinyl migration is favoured over aryl migration by electron-donating substituents on the vinyl ligand, electron-withdrawing substituents on the aryl ligand, good σ-donor phosphorus ligands and use of tBuNC as the incoming ligand. The rate of phenyl migration in Ru(CO)2(CH=CHPh)Ph(PMe2Ph)2 is independent of tBuNC concentration: k=1.5 × 10−3 s−1 at 20°C. Isomer 3 of complexes Ru(CO)2(CH=CHR) (C6H4X-4)L2 in which the phosphorus ligands are mutually trans is much less reactive towards migration reactions. The reactivity of isomer 1 is attributed to the steric strain of two mutually cis phosphorus ligands.  相似文献   

9.
A series of zirconium(IV) complexes, [ZrX2(XDK)], where XDK is the constrained carboxylate ligand m-xylylenediamine bis(Kemp's triacid imide), were prepared and structurally characterized. The solid state structure of the mononuclear carboxylate alkyl complex [Zr(CH2Ph)2(XDK)] reveals that one benzyl group is bonded in an η2-fashion to the metal center. The reactivity of [Zr(CH2Ph)2(XDK)] displays its electrophilic character toward nucleophiles strong enough to displace the η2-benzyl group. Thus, weak sigma donor ligands such as CO, alkynes and anilines do not react, whereas strong sigma donors, such as pyridines and isocyanides, rapidly form the monoadduct [Zr(CH2Ph)2(4-tert-butylpyridine)(XDK)] and [Zr{η2-2,6-Me2PhNCCH2Ph}2(XDK)], an η2-iminoacyl derivative, respectively. Attempts to prepare zirconium amido complexes with H2XDK generally afforded the eight-coordinate [Zr(XDK)2] complex but use of the small amido ligand precursorZr(NMe2)4 allowed [Zr(NMe2)2(4-tert-butylpyridine)(XDK)] to be isolated in good yield.  相似文献   

10.
New tin and thallium reagents capable of transferring the 1,2-di-tert-butylcyclopentadienyl moiety are easily prepared and utilized in furthering the transition metal organometallic chemistry of this intersting ligand. Lithium 1,2-di-tert-butylcyclopentadienide (1) reacts cleanly and selectively with SnClMe3 to give 2,3-di-tert-butyl-5-trimethylstannyl-1,3-cyclopentadiene (2), which in turn reacts with Re(CO)5Br to form the half-sandwich complex [Re(η5-C5H3(1,2-But)2)(CO)3] (3). The reaction between thallium ethoxide and 1,5-ditert-butyl-1,3-cyclopentadiene in hexane affords the excellent cyclopentadienyl transfer reagent, thallium 1,2-di-tert-butylcyclopentadienide (4). The thallium salt reacts with [Ru(COD)Cl2]n to give the sandwich complex [Ru(η5-C5H3(1,2-Bu2t)2)] (5).  相似文献   

11.
The reactivity, towards nucleophiles and electrophiles, of dimolybdenum allenylidene complexes of the type [Cp2Mo2(CO)4(μ,η2(4e)-C=C=CR1R2)] (Cp=η5-C5H5) has been investigated. The nucleophilic attacks occur at the Cγ carbon atom, while electrophiles affec the C atom. Variable temperature solution 1H NMR studies show a dynamic behavior of these complexes consisting of an equilibrium between two enantiomers with a symmetrical [Cp2Mo2(CO)4(μ-σ,σ(2e)-C=C=CR1R2)] transition state. Extended Hückel MO calculations have been carried out on the model [Cp2Mo2(CO)4(μ,η2-C=C=CH2]. The calculated charges of the allenylidene carbon atoms suggest that the electrophilic attacks are under charge control, while the nucleophilic attacks are rather under orbital control.  相似文献   

12.
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   

13.
The reaction of RuCl3(H2O), with C5Me4CF3J in refluxing EtOH gives [Ru25-C5Me1CF2)2 (μ-Cl2] (20 in 44% yield. Dimer 2 antiferromagnetic (−2J=200 cm1). The crystal structures of 2 (rhombohedral system, R3 space group, Z=9, R=0.0589) and [Rh25-C5Me4CF3(2Cl2(μ-Cl)2] (3) (rhombohedral system. space group, Z = 9, R = 0.0641) were solved; both complexes have dimeric structures with a trans arrangement of the η5-C5Me4CF4 rings. Comparison of the geometry of 2 and 3 with those of the corresponding η5-C5Me5 complexes shows that lowering the ring symmetry causes significant distortion of the M2(μ-Cl)2 moiety. The analysis of the MCl3 fragment conformations in 2 and 3 and in the η5-C5ME5 analogues shows that they are correlated with the M---M distances. The Cl atoms are displaced by Br on reaction of 2 with KBr in MeOH to give the diamagnetic dimer [Ru25-C5Me4CF3)2Br2 (μ-Br2] (4). Complex 2 reacts with O2 in CH2Cl2 solution at ambient temperature to form a mixture of isomeric η6-fulvene dimers [Ru26-C5Me3CF3 = CH2)2Cl2(μ-Cl)2] (5). Reactions of 5 with CO and allyl chloride give Ru(η5-C5Me3CF3CH2Cl)(CO)2Cl (6) and Ru(η5-C5Me3CF3CF3CH2Cl)(η3-C3H5)Cl2 (7) respectively.  相似文献   

14.
Kinetic results are reported for intramolecular PPh3 substitution reactions of Mo(CO)21-L)(PPh3)2(SO2) to form Mo(CO)22-L)(PPh3)(SO2) (L = DMPE = (Me)2PC2H4P(Me)2 and dppe=Ph2PC2H4PPh2) in THF solvent, and for intermolecular SO2 substitutions in Mo(CO)32-L)(η2-SO2) (L = 2,2′-bipyridine, dppe) with phosphorus ligands in CH2Cl2 solvent. Activation parameters for intramolecular PPh3 substitution reactions: ΔH values are 12.3 kcal/mol for dmpe and 16.7 kcal/mol for dppe; ΔS values are −30.3 cal/mol K for dmpe and −16.4 cal/mol K for dppe. These results are consistent with an intramolecular associative mechanism. Substitutions of SO2 in MO(CO)32-L)(η2-SO2) complexes proceed by both dissociative and associative mechanisms. The facile associative pathways for the reactions are discussed in terms of the ability of SO2 to accept a pair of electrons from the metal, with its bonding transformations of η2-SO2 to η1-pyramidal SO2, maintaining a stable 18-e count for the complex in its reaction transition state. The structure of Mo(CO)2(dmpe)(PPh3)(SO2) was determined crystallographically: P21/c, A=9.311(1), B = 16.344(2), C = 18.830(2) Å, ß=91.04(1)°, V=2865.1(7) Å3, Z=4, R(F)=3.49%.  相似文献   

15.
Using thermal and photochemical methods a series of new chromium complexes has been prepared: (ν6-p-C6H4F2)Cr(CO)3; (ν6-C6H5CF3)Cr(CO)3; [m-C6H4(CF3)2]Cr(CO)3; (ν6-C6H5F)Cr(CO)2H(SiCl3); (ν6-C6H5F)Cr(CO)2(SiCl3)2; (p-C6H4F2)Cr(CO)2-H(SiCl3); (C6H5CF3)Cr(CO)2H(SiCl3(p-C6H4F2)Cr(CO)2(SiCl3)2; C6H5CF3)Cr(CO)2(SiCl3)2; [m-C6H4(CF3)2]Cr(CO)2-H(SiCl3); [m-C6H4(CF3)2]Cr(CO)2(SiCl3)2. Two compounds were structurally characterized by X-ray diffraction. These data combined with IR and 1H NMR have allowed assessment of some of the electronic and steric effects. The Cr-arene bond is considerably longer in the Cr(II) derivatives than in the Cr(0) species. Also the Cr center, as might be expected, is less electron rich in the Cr(II) dicarbonyl disilyl derivatives. The ν6-p-C6H4F2 ligands are slightly folded so that the C---F carbons are moved further away from the Cr center. Comparison of structural and electronic effects is made with a series of similar chromium compounds reported in the literature. These new (arene)Cr(II) derivatives possess more labile ν6-arene ligands, which promise a rich chemistry at the chromium center.  相似文献   

16.
Carbonylation of the anionic iridium(III) methyl complex, [MeIr(CO)2I3] (1) is an important step in the new iridium-based process for acetic acid manufacture. A model study of the migratory insertion reactions of 1 with P-donor ligands is reported. Complex 1 reacts with phosphites to give neutral acetyl complexes, [Ir(COMe)(CO)I2L2] (L = P(OPh)3 (2), P(OMe)3 (3)). Complex 2 has been isolated and fully characterised from the reaction of Ph4As[MeIr(CO)2I3] with AgBF4 and P(OPh)3; comparison of spectroscopic properties suggests an analogous formulation for 3. IR and 31P NMR spectroscopy indicate initial formation of unstable isomers of 2 which isomerise to the thermodynamic product with trans phosphite ligands. Kinetic measurements for the reactions of 1 with phosphites in CH2Cl2 show first order dependence on [1], only when the reactions are carried out in the presence of excess iodide. The rates exhibit a saturation dependence on [L] and are inhibited by iodide. The reactions are accelerated by addition of alcohols (e.g. 18× enhancement for L = P (OMe)3 in 1:3 MeOH-CH2Cl2). A reaction mechanism is proposed which involves substitution of an iodide ligand by phosphite, prior to migratory CO insertion. The observed rate constants fit well to a rate law derived from this mechanism. Analysis of the kinetic data shows that k1, the rate constant for iodide dissociation, is independent of L, but is increased by a factor of 18 on adding 25% MeOH to CH2Cl2. Activation parameters for the k1 step are ΔH = 71 (±3) kJ mol, ΔS = −81 (±9) J mol−1 K−1 in CH2Cl2 and ΔH = 60(±4) kJ mol−1, ΔS = −93(± 12) J mol−1 K−1 in 1:3 MeOH-CH2Cl2. Solvent assistance of the iodide dissociation step gives the observed rate enhancement in protic solvents. The mechanism is similar to that proposed for the carbonylation of 1.  相似文献   

17.
The reaction of [N(PPh3)2]2[Ni6(CO)12] with Cu(PPh3)xCl (x=1, 2), as well as the degradation of [N(PPh3)2]2[H2Ni12(CO)21] with PPh3, affords the new and unstable dark orange–brown [N(PPh3)2]2[Ni9(CO)16].THF salt in low yields. This salt has been characterized by a CCD X-ray diffraction determination, along with IR spectroscopy and elemental analysis. The close-packed two-layer metal core geometry of the [Ni9(CO)16]2− dianion is directly related to that of the bimetallic [Ni6Rh3(CO)17]3− trianion and may be envisioned to be formally derived from the hcp three-layer geometry of [Ni12(CO)21]4− by the substitution of one of the two outer [Ni3(CO)3(μ−CO)3]2− layers with a face-bridging carbonyl group.  相似文献   

18.
Reactions of [(PPh3)2Pt(η3-CH2CCPh)]OTf with each of PMe3, CO and Br result in the addition of these species to the metal and a change in hapticity of the η3-CH2CCPh to η1-CH2CCPh or η1-C(Ph)=C=CH2. Thus, PMe3 affords [(PMe3)3Pt(η1-C(Ph)=C=CH2)]+, CO gives both [trans-(PPh3)2Pt(CO)(η1-CH2CCPh)]+ and [trans-(PPh3)2Pt(CO)(η1-C(Ph)=C=CH2)]+, and LiBr yields cis-(PPh3)2PtBr(η1-CH2CCPh), which undergoes isomerization to trans-(PPh3)2PtBr(η1-CH2CCPh). Substitution reactions of cis- and trans-(PPh3)2PtBr(η1-CH2CCPh) each lead to tautomerization of η1-CH2CCPh to η1-C(Ph)=C=CH2, with trans-(PPh3)2PtBr(η1-CH2CCPh) affording [(PMe3)3Pt(η1-C(Ph)=C=CH2)]+ at ambient temperature and the slower reacting cis isomer giving [trans-(PPh3)(PMe3)2Pt(η1-C(Ph)=C=CH2)]+ at 54 °C . All new complexes were characterized by a combination of elemental analysis, FAB mas spectrometry and IR and NMR (1H, 13C{1H} and 31P{1H}) spectroscopy. The structure of [(PMe3)3Pt(η1-C(Ph)=C=CH2)]BPh4·0.5MeOH was determined by single-crystal X-ray diffraction analysis.  相似文献   

19.
The Pt2 (II) isomeric terminal hydrides [(CO)(H)Pt(μ-PBu2)2Pt(PBu2H)]CF3SO3 (1a), and [(CO)Pt(μ-PBu2)2Pt(PBu2H)(H)]CF3SO3 (1b), react rapidly with 1 atm of carbon monoxide to give the same mixture of two isomers of the Pt2 (I) dicarbonyl [Pt2(μ-PBu2)(CO)2(PBu2H)2]CF3SO3 (3-Pt); the solid state structure of the isomer bearing the carbonyl ligands pseudo-trans to the bridging phosphide was solved by X-ray diffraction. A remarkable difference was instead found between the reactivity of 1a and 1b towards carbon disulfide or isoprene. In both cases 1b reacts slowly to afford [Pt2(μ-PBu2)(μ,η22-CS2)(PBu2H)2]CF3SO3 (4-Pt), and [Pt2(μ-PBu2)(μ,η22-isoprene) (PBu2H)2]CF3SO3 (6-Pt), respectively. In the same experimental conditions, 1a is totally inert. A common mechanism, proceeding through the preassociation of the incoming ligand followed by the P---H bond formation between one of the bridging P atoms and the hydride ligand, has been suggested for these reactions.  相似文献   

20.
Cis(or trans)-[RuCl2(CO)2(PPh3)2] react with two and one equivalents of AgBF4 to give the recently reported [Ru(CO)2(PPh3)2][BF4]2·CH2Cl2 (1) and novel [RuCl(CO)2(PPh3)2][BF4] · 1/2 CH2Cl2 (2), respectively. Cis-[RuCl2(CO)2(PPh3)2] also reacts with two equivalents of AgBF4 in the presence of CO to give [Ru(CO)3(PPh3)2][BF4]2 (3). Reactions of 1 and 2 with NaOMe and CO at 1 atm produce the carbomethoxy species [Ru(COOMe)2(CO)2(PPh3)2] (4) and [RuCl(COOMe)(CO)2(PPh3)2] (5), respectively. Complex 4 can also be formed from the reaction of 3 with NaOMe and CO. Alternatively, 4 is formed from cis-[RuCl2(CO)2(PPh3)2] with NaOMe and CO at elevated pressure (10 atm); if these reactants are refluxed under 1 atm of CO, [Ru(CO)3(PPh3)2] is the product. The reaction of [RuCl(CO)3(PPh3)2][AlCl4] with NaOMe provides an alternative route to the preparation of 5, but the product is contaminated with [RuCl2(CO)2(PPh3)2]. Compounds 1. 2, 4 and 5 have been characterised by IR, 1H NMR and analysis, whilst the formulation of 3 is proposed from spectroscopic data only. This account also examines the reactivity of [Ru(CO)2(PPh3)2][BF4]2 · CH2Cl2 with NaBH4, conc. HCl, KI and, finally, MeCOONa in the presence of CO. The products of these reactions, namely cis-[RuH2(CO)2(PPh3)2], cis-[RuCl2(CO)2(PPh3)2], cis-[RuI2(CO)2(PPh3)2] and [Ru(OOCMe)2(CO)2(PPh3)2], have been identified by comparison of their spectra with previous literature.  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