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 46 毫秒
1.
Treatment of the 16-electron hydroxy hydride complex [Ru(IMes)2(CO)H(OH)] (1, IMes = 1,3-bis-(2,4,6-trimethylphenyl)imidazol-2-ylidene) with HCCR affords the alkynyl species [Ru(IMes)2(CO)H(CCR)] (R = Ph 3, SiMe3, 4) and [Ru(IMes)2(CO)(CCR)2] (R = Ph, 5). Deuterium labelling studies show that the mono-alkynyl complexes are formed via hydrogen transfer from a coordinated alkyne ligand to Ru-OH, while bis-alkynyl formation is proposed to take place through hydrogen transfer to Ru-H. Both 3 and 5 readily coordinate CO to give the corresponding dicarbonyl species 6 and 7. Addition of HCCPh to the hydride chloride precursor [Ru(IMes)2(CO)HCl] (2) results in a different reaction pathway involving alkyne insertion into the Ru-H bond to yield the alkenyl chloride complex [Ru(IMes)2(CO)(CHCHPh)Cl] 8. Complexes 3-8 have been structurally characterised by X-ray crystallography.  相似文献   

2.
Two mononuclear mixed-ligand ruthenium(III) complexes with oxalate dianion (ox2−) and acetylacetonate ion (2,4-pentanedionate, acac), K2[Ru(ox)2(acac)] (1) and K[Ru(ox)(acac)2] (2), were prepared as a candidate for a building block. In fact, reaction of complex 2 with manganese(II) sulfate gave a heterometallic tetranuclear complex, TBA[MnII{(μ-ox)RuIII(acac)2}3] (5) in the presence of tetrabutylammonium (TBA) bromide. The 1H NMR, UV-Vis, selected IR and FAB mass spectral data of these complexes are presented. Both mixed-ligand ruthenium(III) complexes gave a Nernstian one-electron reduction step in 0.1 mol dm−3 Na2SO4 aqueous solution on a mercury electrode at 25 °C. Comparison of observed reversible half-wave potentials with calculated values for a series of [Ru(ox)n(acac)3 − n]n (n=0-3) complexes by using Lever’s ligand electrochemical parameters is presented.  相似文献   

3.
Reactions of ligands 1-ethyl-5-methyl-3-phenyl-1H-pyrazole (L1) and 5-methyl-1-octyl-3-phenyl-1H-pyrazole (L2) with [PdCl2(CH3CN)2 and K2PtCl4 gave complexes trans-[MCl2(L)2] (L = L1, L2). The new complexes were characterised by elemental analyses, conductivity measurements, infrared, 1H and 13C{1H} NMR spectroscopies and X-ray diffraction. The NMR study of the complex [PdCl2(L1)2], in CDCl3 solution, is consistent with a very slow rotation of ligands around the Pd-N bond, so that two conformational isomers can be observed in solution (syn and anti). Different behaviour is observed for complexes [PdCl2(L2)2] and [PtCl2(L)2] (L = L1, L2), which present an isomer in solution at room temperature (anti). The crystal structure of [PdCl2(L1)2] complex is described, where the Pd(II) presents a square planar geometry with the ligands coordinated in a trans disposition.  相似文献   

4.
Ray K  Lee SM  Que L 《Inorganica chimica acta》2008,361(4):1066-1069
The mechanism of formation of [FeIV(O)(N4Py)]2+ (2, N4Py = N,N-bis(2-pyridylmethyl)-N-bis(2-pyridyl)methylamine) from the reaction of [FeII(N4Py)(CH3CN)]2+ (1) with m-chloroperbenzoic acid (mCPBA) in CH2Cl2 at −30 °C has been studied on the basis of the visible spectral changes observed and the reaction stoichiometry. It is shown that the conversion of 1 to 2 in 90% yield requires 1.5 equiv. peracid and takes place in two successive one-electron steps via an [FeIII(N4Py)OH]2+(3) intermediate. The first oxidation step uses 0.5 equiv. peracid and produces 0.5 equiv. 3-chlorobenzoic acid, while the second step uses 1 equiv. peracid and affords byproducts derived from chlorophenyl radical. We conclude that the FeII(N4Py) center promotes O-O bond heterolysis, while the FeIII(N4Py) center favors O-O bond homolysis, so the nature of O-O bond cleavage is dependent on the iron oxidation state.  相似文献   

5.
Reaction of NiI2 with the PCP-ligand {1-Et-2,6-(CH2PiPr2)2-C6H3} (1) results in selective activation of the strong sp2-sp3 aryl-ethyl bond to afford the aryl-nickel complex [Ni{2,6-(CH2PiPr2)2-C6H3}I] (2), whereas reaction of NiI2 with {1,3,5-(CH3)3-2,6-(CH2PiPr2)2-C6H} (4) leads to the formation of the benzylic complex [Ni{1-CH2-2,6-(CH2PiPr2)2-3,5-(CH3)2-C6H}I] (5) by selective C-H bond activation. Thermolysis of 5 results in formation of [Ni{2,6-(CH2PiPr2)2-3,5-(CH3)2-C6H}I] (6) by activation of the sp2-sp3 C-C bond. The identity of the new 16-electron complexes 2 and 6 was confirmed by reaction of NiI2 with {1,3-(CH2PiPr2)2-C6H4} (3) and {1,3-(CH3)2-4,6-(CH2PiPr2)2-C6H2} (7), respectively, lacking the aryl-alkyl groups between the “phosphines arms” (alkyl=ethyl, methyl). Complexes 2 and 5 have been fully characterized by X-ray analysis. Nickel-based activation of an unstrained C-O single bond was observed as well. Reaction of the aryl-methoxy bisphosphine {1-OMe-2,6-(CH2PiPr2)2-C6H3} (8) with NiI2 results in the formation of the phenoxy complex [Ni{1-O-2,6-(CH2PiPr2)2-C6H3}I] (9) by selective sp3-sp3 C-O bond activation.  相似文献   

6.
A synthetic and mechanistic study is reported on ligand substitution and other reactions of six-coordinate ruthenium(II) carbonyl complexes containing tridentate PhP(CH2CH2CH2PCy2)2 (Cyttp). Carbonylation of cis-mer-Ru(OSO2CF3)2(CO)(Cyttp) (1) affords [cis-mer-Ru(OSO2CF3)(CO)2(Cyttp)]O3SCF3 (2(O3SCF3)) and, on longer reaction times, [cis-mer-Ru(solvent)(CO)2(Cyttp)](O3SCF3)2 (solvent = acetone, THF, methanol). 2(O3SCF3) reacts with each of NaF, LiCl, LiBr, NaI, and LiHBEt3 to yield [cis-mer-RuX(CO)2(Cyttp)]+ (X = F (3), Cl (4), Br (5), I (6), H (7)), isolated as 3-7(BPh4). These conversions proceed with high stereospecificity to afford only a single isomer of the product that is assigned a structure in which the Ph group of Cyttp points toward the CO trans to X (anti when X = F, Cl, Br, or I; syn when X = H). Treatment of 2(O3SCF3) with NaOMe and CO generates the methoxycarbonyl complex [cis-mer-Ru(CO2Me)(CO)2(Cyttp)]+ (8), whereas addition of excess n-BuLi to 2(O3SCF3) in THF under CO affords mer-Ru(CO)2(Cyttp) (9). The two 13C isotopomers [cis-mer-Ru(OSO2CF3)(CO)(13CO)(Cyttp)]O3SCF3 (2′(O3SCF3): 13CO trans to PC; 2″(O3SCF3): 13CO cis to all P donors) were synthesized by appropriate adaptations of known transformations and used in mechanistic studies of reactions with each of LiHBEt3, NaOMe/CO, and n-BuLi. Whereas LiHBEt3 reacts with 2′(O3SCF3) and 2″(O3SCF3) to replace triflate by hydride without any scrambling of the carbonyl ligands, the corresponding reactions of NaOMe-CO are more complex. The methoxide combines with the CO cis to triflate in 2, and the resultant methoxycarbonyl ligand ends up positioned trans to the incoming CO in 8. A mechanism is proposed for this transformation. Finally, treatment of either 2′(O3SCF3) or 2″(O3SCF3) with an excess of n-BuLi leads to the formation of the same two ruthenium(0) isomers of mer-Ru(CO)(13CO)(Cyttp). These products represent, to our knowledge, the first example of a syn-anti pair of isomers of a five-coordinate metal complex.  相似文献   

7.
The reactions of the N-heterocyclic carbene (NHC) stabilised group 13 trihydride complexes [AlH3(IMeMe)] (1) (IMeMe = 1,3,4,5-tetramethylimidazol-2-ylidene), [AlH3(IiPrMe)] (2) (IiPrMe = 1,3-diisopropyl-4,5-dimethylimidazol-2-ylidene) with three molar equivalents of phenol, and [InH3(IMes)] (3) (IMes = 1,3-bis(2,4,6-trimethylphenyl)imidazole-2-ylidene) with one molar equivalent of 1,1,1,5,5,5-hexafluoropentan-2,4-dione (F6acacH) are presented. These render the imidazolium tetraphenoxyaluminate species; [IMeMe · H][Al(OPh)4] (4) and [IiPrMe · H][Al(OPh)4] (5), and 1,3-bis(2,4,6-trimethylphenyl)imidazolium 1,1,1,5,5,5-hexafluoropentan-2,4-dionate; [IMes · H][CH{C(O)CF3}2] (6), the latter leading to metallohydride decomposition. The molecular structures of 4 and 6 are described.  相似文献   

8.
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   

9.
The reaction between the dirhenium(III,III) anion, [Re2Cl8]2−, and the secondary phosphine, PCy2H, yields a mixture of products as a result of disproportionation, namely, a dirhenium(II,III) chloride-phosphine complex 1,3,6-Re2Cl5(PCy2H)3 (1) and a dirhenium(IV) face-sharing bioctahedral compound with bridging phosphido groups, [Bu4N][Re2(μ-PCy2)3Cl6] (2). The diphenylphosphine analogue of 2, [Bu4N][Re2(μ-PPh2)3Cl6] (3) has been similarly prepared from the reaction of [Re2Cl8]2− with PPh2H. An interesting dirhenium(III,III) complex, [Bu4N]2[Re2(μ-PPh2)2(PPh2H)2Cl6] (4) having both neutral terminal phosphines and anionic phosphido bridges, has also been isolated as an intermediate in the latter system. Crystal structures of 1-4 have been determined by X-ray crystallography. The compounds were also characterized by cyclic voltammetry, IR and 31P NMR spectroscopy.  相似文献   

10.
The 16-electron, coordinatively unsaturated, dicationic ruthenium complex [Ru(P(OH)2(OMe))(dppe)2][OTf]2 (1a) brings about the heterolysis of the C-H bond in phenylacetylene to afford the phenylacetylide complex trans-[Ru(CCPh)(P(OH)2(OMe))(dppe)2][OTf] (2). The phenylacetylide complex undergoes hydrogenation to give a ruthenium hydride complex trans-[Ru(H)(P(OH)2(OMe))(dppe)2][OTf] (3) and phenylacetylene via the addition of H2 across the Ru-C bond. The 16-electron complex also reacts with HSiCl3 quite vigorously to yield a chloride complex trans-[Ru(Cl)(P(OH)2(OMe))(dppe)2][OTf] (4). On the other hand, the other coordinatively unsaturated ruthenium complex [Ru(P(OH)3)(dppe)2][OTf]2 (1b) reacts with a base N-benzylideneaniline to afford a phosphonate complex [Ru(P(O)(OH)2)(dppe)2][OTf] (5) via the abstraction of one of the protons of the P(OH)3 ligand by the base. The phenylacetylide, chloride, and the phosphonate complexes have been structurally characterized. The phosphonate complex reacts with H2 to afford the corresponding dihydrogen complex trans-[Ru(η2-H2)(P(O)(OH)2)(dppe)2][OTf] (5-H2). The intact nature of the H-H bond in this species was established using variable temperature 1H spin-lattice relaxation time measurements and the observation of a significant J(H,D) coupling in the HD isotopomer trans-[Ru(η2-HD)(P(O)(OH)2)(dppe)2][OTf] (5-HD).  相似文献   

11.
New five mono- and dinuclear Ir hydrido complexes with polydentate nitrogen ligands, [Ir(H)2(PPh3)2(tptz)]PF6 (1), [Ir2(H)4(PPh3)4(tptz)](PF6)2 · 2H2O (2 · 2H2O), [Ir(H)2(PPh3)2(tppz)]BF4 (3), [Ir2(H)4(PPh3)4(tppz)](BF4)2 (4) and [Ir2(H)4(PPh3)4(bted)](BF4)2 · 6CHCl3 (5 · 6CHCl3), were systematically prepared by the reactions of the precursor Ir hydrido complex [Ir(H)2(PPh3)2(Me2CO)2]X (X=PF6 and BF4) with 2,4,6-tris(2-pyridyl)-1,3,5-triazine (tptz), 2,3,5,6-tetrakis(2-pyridyl)pyrazine (tppz) and 1,4-bis(2,2:6,2″-terpyridine-4-yl)benzene (bted), and their structures and properties were characterized in the solid state and in solution. Each of the Ir hydrido complexes with polydentate nitrogen ligands crystallographically described a unique coordination mode. Their 1H NMR spectra demonstrated unusual 1H NMR chemical shifts of pyridyl rings that are likely induced by the ring current effect of neighboring ligands.  相似文献   

12.
A series of cationic, half-sandwich ruthenium complexes with the general formula [(η6-p-cymene)RuCl(MeSC6H42-NCHAr)][PF6] (3a-h), have been prepared from the reaction of [(η6-p-cymene)RuCl2]2 with various N,S-donor Schiff base ligands derived from 2-(methylthio)aniline and several substituted benzaldehydes. The related aniline complex [(η6-p-cymene)RuCl(MeS-C6H4-2-NH2)][PF6] (4) was synthesized from 2-(methylthio)aniline. All of the ruthenium complexes were characterized by IR, 1H NMR, and UV/Vis spectroscopies. The molecular structure of complex 4 was determined by X-ray crystallography.  相似文献   

13.
A series of new iridium(III) complexes containing pentamethylcyclopentadienyl (Cp = η5-C5Me5) and 1,8-naphthyridine (napy) have been prepared. X-ray crystallography revealed that napy acted as a monodentate, a didentate chelating, and a bridging ligand in complexes of [CpIrCl2(napy)] (1), [CpIrCl(napy)]PF6 (2), and [(CpIrCl)2(H)(napy)]PF6 (4), respectively. The crystal structure of [CpIr(napy)2](PF6)2 (3) has also been determined; the dicationic complex bore both monodentate and chelating napy ligands. Dinuclear CpIrIII complex bridged by napy was only isolable if two IrIII centers were supported by a hydride (H) bridge. In complexes 2 and 3, the four-membered chelate rings formed by napy exhibited a large steric strain; in the rings the NIrN bond angles were only 60.5(2)-61.0(4)° and the IrNC angles were 94.7(8)-96.7(8)°. The bridging coordination of napy in complex 4 also afforded a large strain, i.e., the IrIII centers were displaced by 0.84(3) Å from the napy plane, due to the steric interaction between two CpIrCl moieties. The monodentate napy complex 1 in CDCl3 or CD2Cl2 at ambient temperature showed a rapid coordination-site exchange reaction, which gave two N sites of napy equivalent; at temperatures below −40 °C, the 1H NMR spectra corresponded to the molecular structure of [CpIrCl2(napy-κN)]. The analogous diazido complex of [CpIr(N3)2(napy)] (5) has also been prepared, and the crystal structure has been determined. In contrast to the dichloro complex 1, the diazido complex 5 exhibited a dissociation equilibrium of coordinated napy in solution.  相似文献   

14.
cis,trans-Fe(CO)2(PMe3)2(p-Y-C6H4)X [X=Br, Y=H (4a), MeO (4b), Cl (4c), F (4d), Me (4e); X=I, Y=H (5); X=Cl, Y=H (6)] and cis,trans-Fe(CO)2(PMe3)2(σ-CHCH2)X [X=Br (7); X=I (8); X=Cl (9)] are prepared by reacting dihalide complexes cis,trans,cis- Fe(CO)2(PMe3)2X2 [X=Br (1), X=I (2), X=Cl (3)] with Grignard reagents p-Y-C6H4-MgBr (Y=H, OMe, Cl, F, Me) or CH2CH-MgBr and with lithium reagents PhLi, CH2CH-Li. With both reagents, the reaction proceeds following two parallel pathways: one is the metallation reaction which yields alkyl derivatives, the other affords 17 electron complexes [Fe(CO)2(PMe3)2X] via monoelectron reductive elimination. The influence of the halides and organometallic reagents on the yield of the metallation reaction is discussed. The solution structure of the complexes is assigned on the basis of IR and 1H, 13C, 19F, 31P NMR spectra. The solid state structure of complexes 4a, 5 and 6 is determined by single crystal X-ray diffractometric methods.  相似文献   

15.
The 2-methallyl complex [(η5-C9H7)Ru(η3-2-MeC3H4)(PPh3)] (3), prepared from [(η5-C9H7)Ru(PPh3)2Cl] (2) and 2-MeC3H4MgCl, reacts with HX (X = Cl, CF3CO2) in the presence of ethene to give the chiral-at-metal compounds [(η5-C9H7)Ru(C2H4)(PPh3)X] (4, 5) in nearly quantitative yields. Treatment of 2 with AgPF6 and ethene affords [(η5-C9H7)Ru(C2H4)(PPh3)2]PF6 (6), which reacts with acetone to give the substitution product [(η5-C9H7)Ru(OCMe2)(PPh3)2]PF6 (7). The molecular structure of 7 has been determined crystallographically. Whereas treatment of 4 with CH(CO2Et)N2 yields the olefin complex [(η5-C9H7)Ru{η2-(Z)-C2H2(CO2Et)2}(PPh3)Cl] (8), the reactions of 4 and 5 with Ph2CN2, PhCHN2 and (Me3Si)CHN2 lead to the formation of the carbeneruthenium(II) derivatives [(η5-C9H7)Ru(CRR′)(PPh3)Cl] (9-11) and [(η5-C9H7)Ru(CRR′)(PPh3)(κ1-O2CCF3)] (12-14), respectively. Treatment of 9 (R = R′ = Ph), 10 (R = H, R′ = Ph) and 11 (R = H, R′ = SiMe3) with MeLi produces the hydrido(olefin) complexes [(η5-C9H7)RuH(η2-CH2CPh2)(PPh3)] (15), [(η5-C9H7)RuH(η2-CH2CHPh)(PPh3)] (18a,b) and [(η5-C9H7)RuH(η2-CH2CHSiMe3)(PPh3)] (19) via C-C coupling and β-hydride shift. The analogous reactions of 11 with PhLi gives the η3-benzyl compound [(η5-C9H7)Ru{η3-(Me3Si)CHC6H5}(PPh3)] (20). The η3-allyl complex [(η5-C9H7)Ru(η3-1-PhC3H4)(PPh3)] (17) was prepared from 10 and CH2CHMgBr by nucleophilic attack.  相似文献   

16.
Substituted salicylaldehydes [C6HR1R2R3(CHO)(OH)] react with CoMe3(PMe3)3 to afford 6-coordinate (cis-dimethyl)(2-formyl-phenolato)trans-bis(trimethylphosphine)cobalt(III) compounds Co[C6HR1R2R3(CHO)(O)Me2](PMe3)2 (1: R1 = H; R2 = Me; R3 = tert-Bu; 2: R1, R2 = C6H4; R3 = H). Accordingly, substituted enolated malonic dialdehydes (CHO-CR4CR5-OH) react with CoMe3(PMe3)3 to afford 6-coordinate (cis-dimethyl)(2-formyl-enolato)trans-bis(trimethylphosphine)cobalt(III) compounds Co[(CHO-CR4CR5-O)(Me)2](PMe3)2 (3: R4, R5 = (CH2)2C6H4; 4: R4 = R5 = C6H5). In the molecular structure of 4, the cobalt atom is centred in an octahedral coordination geometry brought about by a six-membered chelate ring (O:O-ligand), cis-dimethyl and trans-trimethylphosphine groups. A reaction mechanism is suggested.  相似文献   

17.
The synthesis and X-ray characterization of binuclear dipalladium(I) and diplatinum(I) p-xylene complexes [Pd26-C8H10)2(μ-Cl/Br)2(GaCl3)2] (1) and [Pt26-C8H10)2(Ga2Br7)2] (5) are reported. It was established that the toluene ligands in the palladium complex [Pd26-C7H8)2(GaCl4)2] (3) can be substituted by naphthalene without disruption of the metal-metal bond. The reaction of 3 with Pd(PPh3)4 leads to the formation of a dipalladium(II) μ-diphenylphosphido compound [Pd2(μ-PPh2)(PPh3)4] (GaCl4)2 · 4(C7H8) (4), most likely also involving a bridging μ-H ligand.  相似文献   

18.
The alkoxo-bridged dinuclear copper(II) complexes [Cu2(ap)2(NO2)2] (1), [Cu2(ap)2(C6H5COO)2] (2) and [Cu2(ap)2μ-1,3-C6H4(COO)2(dmso)2]·dmso (3) (ap = 3-aminopropanolato and dmso = dimethyl sulfoxide) have been synthesized via self-assembly from copper(II) perchlorate, 3-aminopropanol as main chelating ligand and nitrite and isophthalate anions as spacers and benzoate anion as auxiliary ligand. Complexes 1 and 3 crystallize as 2D and 1D coordination polymers, respectively, and their structures consist of dinuclear [Cu2(ap)2]2+ units connected with nitrite and isophthalate ligands. The adjacent dinuclear units of 2 and 1D polymers of 3 are further connected by hydrogen bonds resulting in the formation of 2D layers. The variable temperature crystallographic measurements of 1 at 100, 173 and 293 K indicate the static Jahn-Teller distortion with librational disorder in the nitrite group. Experimental magnetic studies showed that complexes 1-3 exhibit strong antiferromagnetic couplings. The values of the magnetic exchange coupling constant for 1-3 are well reproduced by the theoretical calculations.  相似文献   

19.
The new complex, [RuII(bpy)2(4-HCOO-4′-pyCH2 NHCO-bpy)](PF6)2 · 3H2O (1), where 4-HCOO-4′-pyCH2NHCO-bpy is 4-(carboxylic acid)-4′-pyrid-2-ylmethylamido-2,2′-bipyridine, has been synthesised from [Ru(bpy)2(H2dcbpy)](PF6)2 (H2dcbpy is 4,4′-(dicarboxylic acid)-2,2′-bipyridine) and characterised by elemental analysis and spectroscopic methods. An X-ray crystal structure determination of the trihydrate of the [Ru(bpy)2(H2dcbpy)](PF6)2 precursor is reported, since it represented a different solvate to an existing structure. The structure shows a distorted octahedral arrangement of the ligands around the ruthenium(II) centre and is consistent with the carboxyl groups being protonated. A comparative study of the electrochemical and photophysical properties of [RuII(bpy)2(4-HCOO-4′-pyCH2NHCO-bpy)]2+ (1), [Ru(bpy)2(H2dcbpy)]2+ (2), [Ru(bpy)3]2+ (3), [Ru(bpy)2Cl2] (4) and [Ru(bpy)2Cl2]+ (5) was then undertaken to determine their variation upon changing the ligands occupying two of the six ruthenium(II) coordination sites. The ruthenium(II) complexes exhibit intense ligand centred (LC) transition bands in the UV region, and broad MLCT bands in the visible region. The ruthenium(III) complex, 5, displayed overlapping LC bands in the UV region and a LMCT band in the visible. 1, 2 and 3 were found, via cyclic voltammetry at a glassy carbon electrode, to exhibit very positive reversible formal potentials of 996, 992 and 893 mV (versus Fc/Fc+) respectively for the Ru(III)/Ru(II) half-cell reaction. As expected the reversible potential derived from oxidation of 4 (−77 mV (versus Fc/Fc+)) was in excellent agreement with that found via reduction of 5 (−84 mV (versus Fc/Fc+)). Spectroelectrochemical experiments in an optically transparent thin-layer electrochemical cell configuration allowed UV-Vis spectra of the Ru(III) redox state to be obtained for 1, 2, 3 and 4 and also confirmed that 5 was the product of oxidative bulk electrolysis of 4. These spectrochemical measurements also confirmed that the oxidation of all Ru(II) complexes and reduction of the corresponding Ru(III) complex are fully reversible in both the chemical and electrochemical senses.  相似文献   

20.
《Inorganica chimica acta》2004,357(9):2483-2493
Hexaaquarhodium(III) perchlorate has been found to be an interesting starting material for the preparation of rhodium compounds. New mononuclear hydrido- and carbonylrhodium compounds with bis(benzimidazol-2-ylmethyl)methylamine (L) and PPh3 of the type [Rh(H)L(PPh3)2](ClO4)2 (1) and [Rh(CO)L(PPh3)2](ClO4) (2) have been synthesized. The monomeric Rh(II) compound was identified as an intermediate. Complexes 1 and 2 were characterized by elemental analysis, mass spectrometry and IR, and NMR spectroscopies. The fluxional behaviour of 2 was studied by variable-temperature 1H and 31P{H} NMR experiments.  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