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.
The syntheses, structures, spectroscopy, and electrochemistry for six Ir(III) and Rh(III) mixed sandwich mononuclear complexes involving tridentate macrocycles and pentamethylcyclopentadienide (Cp*) are reported. The complexes are readily prepared by direct ligand substitution reactions from the dichloro bridged binuclear complexes, [{M(Cp*)(Cl)2}2]. All complexes have the general formula [M(L)(Cp*)]X2 (M = Ir(III) or Rh(III), L = macrocycle, or Cl) and exhibit a distorted octahedral structure involving three donor atoms from the macrocycle and the facially coordinating carbocyclic Cp* ligand. The complex cations include: [Rh(η5 -Cp*)(9S3)]2+ (1), [Rh(η5-Cp*)(9N3)]2+ (2), [Rh(η5-Cp*)(10S3)]2+ (3), [Ir(η5-Cp*)(9S3)]2+ (4), [Ir(η5-Cp*)(9N3)]2+ (5), and [Ir(η5-Cp*)(10S3)]2+ (6), where 9S3 = 1,4,7-trithiacyclononane, 9N3 = 1,4,7-triazacyclononane, and 10S3 = 1,4,7-trithiacyclodecane. The structures for all six complexes are supported by 1H and 13C{1H} NMR spectroscopy, and five complexes are also characterized by single-crystal X-ray crystallography (complexes 1-5). The 1H NMR splittings between the two sets of methylene protons for both the Rh(III) and Ir(III) 9S3 complexes are much larger (0.4 vs. 0.2 ppm) compared to those in the two 9N3 complexes. Similarly, the 13C{1H} NMR spectra in all four thioether complexes show that the ring carbons in the Cp* ligand are shifted by over 10 ppm downfield compared to the azacrown complexes. The electrochemistry of the complexes is surprisingly invariable and is dominated by a single irreversible metal-centered reduction near −1.2 V vs. Fc/Fc+.  相似文献   

2.
Reactions of [Cp*M(μ-Cl)Cl]2 (M = Ir, Rh; Cp* = η5-pentamethylcyclopentadienyl) with bi- or tri-dentate organochalcogen ligands Mbit (L1), Mbpit (L2), Mbbit (L3) and [TmMe] (L4) (Mbit = 1,1′-methylenebis(3-methyl-imidazole-2-thione); Mbpit = 1,1′-methylene bis (3-iso-propyl-imidazole-2-thione), Mbbit = 1,1′-methylene bis (3-tert-butyl-imidazole-2-thione)) and [TmMe] (TmMe = tris (2-mercapto-1-methylimidazolyl) borate) result in the formation of the 18-electron half-sandwich complexes [Cp*M(Mbit)Cl]Cl (M = Ir, 1a; M = Rh, 1b), [Cp*M(Mbpit)Cl]Cl (M = Ir, 2a; M = Rh, 2b), [Cp*M(Mbbit)Cl]Cl (M = Ir, 3a; M = Rh, 3b) and [Cp*M(TmMe)]Cl (M = Ir, 4a; M = Rh, 4b), respectively. All complexes have been characterized by elemental analysis, NMR and IR spectra. The molecular structures of 1a, 2b and 4a have been determined by X-ray crystallography.  相似文献   

3.
The polyfunctional (H)PNX (X = O or N) ligands 1 and 2 react with [Rh(CO)2Cl]2 to give the corresponding chloro carbonyl complexes {Rh[κ2-(H)PN](CO)Cl} (1a and 2a), where the neutral ligands coordinate in a κ2-PN bidentate fashion, the square planar coordination being completed by the CO trans to N and the chloride trans to P. In chloroform solution 1a maintains its original structure, while 2a partially transforms into the cationic species {Rh[κ3-(H)PNO](CO)}Cl. The chloroform solutions of 1a and 2a react with AgPF6 to give the purely cationic species {Rh[κ3-(H)PNO](CO)}PF6 ([1a]+ and [2a]+), while addition of Et3N originates the neutral species {Rh[κ3-PNN′](CO)} (1b and 2b). All the complexes have been characterized by microanalysis, IR, 1H NMR as well as 31P{1H} NMR spectroscopy. The X-ray structures of ligand 1 and complex 1b are also reported.  相似文献   

4.
The complex [Rh(CO)2Cl]2 reacts with two molar equivalent of pyridine carboxylic acids ligands Py-2-COOH(a), Py-3-COOH(b) and Py-4-COOH(c) to yield rhodium(I) dicarbonyl chelate complex [Rh(CO)2(L/)](1a) {L/ = η2-(N,O) coordinated Py-2-COO(a/)} and non-chelate complexes [Rh(CO)2ClL//](1b,c) {L// = η1-(N) coordinated Py-3-COOH(b), Py-4-COOH(c)}. The complexes 1 undergo oxidative addition (OA) reactions with different electrophiles such as CH3I, C2H5I, C6H5CH2Cl and I2 to give penta coordinated Rh(III) complexes of the types [Rh(CO)(CORn)XL/], {n = 1,2,3; R1 = CH3(2a); R2 = C2H5(3a); X = I and R3 = CH2C6H5 (4a); X = Cl}, [Rh(CO)I2L/](5a), [Rh(CO)(CORn)ClXL//] {R1 = CH3(6b,c); R2 = C2H5(7b,c); X = I and R3 = CH2C6H5 (8b,c); X = Cl} and [Rh(CO)ClI2L//](9b,c). The complexes have been characterized by elemental analysis, IR and 1H NMR spectroscopy. Kinetic data for the reaction of 1a–b with CH3I indicate a first order reaction. The catalytic activity of 1a–c for the carbonylation of methanol to acetic acid and its ester is evaluated and a higher turn over number (TON = 810–1094) is obtained compared with that of the well-known commercial species [Rh(CO)2I2] (TON = 653) at mild reaction conditions (temperature 130 ± 5 °C, pressure 35 ± 5 bar).  相似文献   

5.
Reactions of CpRuCl(PPh3)2 with bis(phosphino)amines, X2PN(R)PX2 (1 R=H, X=Ph; 2 R=X=Ph; 3 R=Ph, X2=O2C6H4) give neutral or cationic mononuclear complexes depending on the reaction conditions. Reaction of 1 with CpRuCl(PPh3)2 gives one neutral complex, [CpRu(Cl)(η2-Ph2PN(H)PPh2)] (4) and two cationic complexes, [CpRu(η2-Ph2PN(H)PPh2)(η1-Ph2PN(H)PPh2)]Cl (5) and [CpRu(PPh3)(η2-Ph2PN(H)PPh2)]Cl (6), whereas the reaction of 2 with CpRuCl(PPh3)2 leads only to the isolation of cationic complex, [CpRu(PPh3)(η2-Ph2PN(Ph)PPh2)]Cl (7). The catechol derivative 3, in a similar reaction, affords an interesting mononuclear complex [CpRu(PPh3){η1-(C6H4O2)PN(Ph)P(O2H4C6)}2]Cl (8) containing two monodentate bis(phosphino)amine ligands. The structural elucidation of the complexes was carried out by elemental analyses, IR and NMR spectroscopic data.  相似文献   

6.
Half-titanocene is well-known as an excellent catalyst for the preparation of SPS (syndiotactic polystyrene) when activated with methylaluminoxane (MAO). Dinuclear half-sandwich complexes of titanium bearing a xylene bridge, (TiCl2L)2{(μ-η5, η5-C5H4-ortho-(CH2–C6H4–CH2)C5H4}, (4 (L = Cl), 7 (L = O-2,6-iPr2C6H3)) and (TiCl2L)2{(μ-η5, η5-C5H4-meta-(CH2–C6H4–CH2)C5H4} (5 (L = Cl), 8(L = O-2,6-iPr2C6H3)), have been successfully synthesized and introduced for styrene polymerization. The catalysts were characterized by 1H- and 13C NMR, and elemental analysis. These catalysts were found to be effective in forming SPS in combination with MAO. The activities of the catalysts with rigid ortho- and meta-xylene bridges were higher than those of catalysts with flexible pentamethylene bridges. The catalytic activity of four dinuclear half-titanocenes increased in the order of 4 < 5 < 7 < 8. This result displays that the meta-xylene bridged catalyst is more active than the ortho-xylene bridged and that the aryloxo group at the titanium center is more effective at promoting catalyst activity compared to the chloride group at the titanium center. Temperature and ratio of [Al]:[Ti] had significant effects on catalytic activity. Polymerizations were conducted at three different temperatures (25, 40, and 70 °C) with variation in the [Al]:[Ti] ratio from 2000 to 4000. It was observed that activity of the catalysts increased with increasing temperature, as well as higher [Al]:[Ti]. Different xylene linkage patterns (ortho and meta) were recognized to be a principal factor leading to the characteristics of the dinuclear catalyst due to its different spatial arrangement, causing dissimilar intramolecular interactions between the two active sites.  相似文献   

7.
The reaction of dimeric rhodium precursor [Rh(CO)2Cl]2 with two molar equivalent of 1,1,1-tris(diphenylphosphinomethyl)ethane trichalcogenide ligands, [CH3C(CH2P(X)Ph2)3](L), where X = O(a), S(b) and Se(c) affords the complexes of the type [Rh(CO)2Cl(L)] (1a–1c). The complexes 1a–1c have been characterized by elemental analyses, mass spectrometry, IR and NMR (1H, 31P and 13C) spectroscopy and the ligands a–c are structurally determined by single crystal X-ray diffraction. 1a–1c undergo oxidative addition (OA) reactions with different electrophiles such as CH3I, C2H5I and C6H5CH2Cl to give Rh(III) complexes of the types [Rh(CO)(COR)ClXL] {R = –CH3 (2a–2c), –C2H5 (3a–3c); X = I and R = –CH2C6H5 (4a–4c); X = Cl}. Kinetic data for the reaction of a–c with CH3I indicate a first-order reaction. The catalytic activity of 1a–1c for the carbonylation of methanol to acetic acid and its ester is evaluated and a higher turn over number (TON = 1564–1723) is obtained compared to that of the well-known commercial species [Rh(CO)2I2] (TON = 1000) under the reaction conditions: temperature 130 ± 2 °C, pressure 30 ± 2 bar and time 1 h.  相似文献   

8.
The mononuclear complexes [(η6-arene)Ru(ata)Cl]PF6 {ata = 2-acetylthiazole azine; arene = C6H6 [(1)PF6]; p-iPrC6H4Me [(2)PF6]; C6Me6 [(3)PF6]}, [(η5-C5Me5)M(ata)]PF6 {M = Rh [(4)PF6]; Ir [(5)PF6]} and [(η5-Cp)Ru(PPh3)2Cl] {η5-Cp = η5-C5H5 [(6)PF6]; η5-C5Me5 (Cp*) [(7)PF6]; η5-C9H7 (indenyl); [(8)PF6]} have been synthesised from the reaction of 2-acetylthiazole azine (ata) and the corresponding dimers [(η6-arene)Ru(μ-Cl)Cl]2, [(η5-C5Me5)M(μ-Cl)Cl]2, and [(η5-Cp)Ru(PPh3)2Cl], respectively. In addition to these complexes a hydrolysed product (9)PF6, was isolated from complex (4)PF6 in the process of crystallization. All these complexes are isolated as hexafluorophosphate salts and characterized by IR, NMR, mass spectrometry and UV–Vis spectroscopy. The molecular structures of [2]PF6 and [9]PF6 have been established by single-crystal X-ray structure analyses.  相似文献   

9.
Reaction of cis-[Mo(NCMe)2(CO)2(η5-L)][BF4] (L=C5H5 or C5Me5) with 1-acetoxybuta-1,3-diene gives the cationic complexes [Mo{η4-syn-s-cis-CH2CHCHCH(OAc)}(CO)2(η5-L)][BF4], which, on reaction with aqueous NaHCO3/CH2Cl2, afford good yields of the anti-aldehyde substituted complexes [Mo{η3-exo-anti-CH2CHCH(CHO)}(CO)2(η5-L)] 2 (L=C5Me5), 4 (L=C5H5)]. The corresponding η5-indenyl substituted complex 5 was prepared by protonation (HBF4·OEt2) of [Mo(η3-C3H5)(CO)2(η5-C9H7)] followed by addition of CH2=CHCH=CH(OAc) and hydrolysis (aq. NaHCO3/CH2Cl2). An X-ray crystallographic study of complex 2 confirmed the structure and showed that there is a contribution from a zwitterionic form involving donation of electron density from the molybdenum to the aldehyde carbonyl group. Treatment of 2 and 4, in methanol solution, with NaBH4 afforded the alcohols [Mo{η3-exo-anti-CH2CHCHCH2(OH)}(CO)2(η5-L)] [6 (L=C5H5), 8 (L=C5Me5)]; however, prolonged (30 h) reaction with NaBH4/MeOH surprisingly gave good yields of the methoxy-substituted complexes [Mo{η3-exo-anti-CH2CHCHCH2(OMe)}(CO)2(η5-L)] [7 (L=C5H5), 9 (L=C5Me5)], the structure of 7 being confirmed by single crystal X-ray crystallography. This methoxylation reaction can be explained by coordination of the hydroxyl group present in 6 and 8 onto B2H6 to form the potential leaving group HOBH3, which on ionisation affords [Mo(η4-exo-buta-1-3-diene)(CO)2(η5-L)]+ which is captured by reaction with OMe. Complex 8 is also formed in good yield on reaction of 2 with HBF4·OEt2 followed by treatment of the resulting cation [Mo{η4-exo-s-cis-syn-CH2CHCHCH(OH)}(CO)2(η5-C5Me5)][BF4] with Na[BH3CN]. Reaction of 4 with the Grignard reagents MeMgI, EtMgBr or PhMgCl afforded moderate yields of the alcohols [Mo{η3-exo-anti-CH2CHCHCH(OH)R}(CO)2(η5-C5H5)] [11 (R=Me), 12 (R=Et), 13 (R=Ph)]. Similarly, treatment of 2 with MeLi gave the corresponding alcohol 14. An attempt to carry out the Oppenauer oxidation [Al(OPr′)3/Me2CO] of 11 resulted in an elimination reaction and the formation of the η3-s-pentadienyl complex [Mo{η3-exo-anti-CH2CHCH(CHCH2)}(CO)2(η5-C5H5)], which was structurally identified by X-ray crystallography. Interestingly, oxidation of 6 with [Bu4nN][RuO4]/morpholine-N-oxide affords the aldehyde complex, 4 in good yield. Finally, reaction of 11 with [NO][BF4] followed by addition of Na2CO3 affords the fur-3-ene complex [Mo{η2-
(H)Me}(CO)(NO)(η5-C5H5)].  相似文献   

10.
The article describes the synthesis and single-crystal X-ray analysis of two sulfato and one thiocyanato copper(II) complex with 2-acetylpyridine S-methylisothiosemicarbazone (HL) of the formulae [Cu(HL)SO4(H2O)]·H2O (1), [Cu2(HL)2(μ-SO4)2]·2H2O (2) and [Cu(HL)(NCS)(SCN)] (3), as well as the structure of the protonated ligand H2L+I. Complexes 1 and 2 were obtained from the reaction of aqueous/methanolic CuSO4·5H2O and ethanolic/methanolic H2L+I solutions, respectively. Complex 3 was synthesized by the reaction of methanolic solutions of Cu(ClO4)2·6H2O, the ligand and NH4SCN, with the addition of triethyl orthoformate. All three complexes have a slightly deformed square-pyramidal structure (τav = 0.15) with the tridentate NNN neutral ligand in the basal plane. In complexes 1 and 3 the apical position is occupied by the oxygen atom of the monodentate SO4 group, or the sulfur atom of the SCN group. Thanks to the hydrogen bonds, complex 3 may be thought of as having a pseudo-dimeric structure. In the authentic centrosymmetric dimer 2, the oxygen atoms of both SO4 groups occupy also the apical position of both coordination polyhedra, as well as an equatorial position. Complexes 1 and 3 have μeff values characteristic of magnetically isolated mononuclear Cu(II) complexes. In contrast to them, complex 2 has a μeff value of 1.57 BM, which is in agreement with its dinuclear structure. All the complexes, in addition to the X-ray analysis and magnetic measurements, were characterized by IR and UV–Vis spectroscopy and by thermal analysis.  相似文献   

11.
A series of neutral pyridine-based organochalcogen ligands, 2,6-bis(1-methylimidazole-2-thione)pyridine (Bmtp), 2,6-bis(1-isopropylimidazole-2-thione)pyridine (Bptp), and 2,6-bis(1-tert-butylimidazole-2-thione)pyridine (Bbtp) have been synthesized and characterized. Reactions of [Cp*M(μ-Cl)Cl]2 (Cp* = η5-pentamethylcyclopentadienyl, M = Ir, Rh) with three pyridine-based organochalcogen ligands result in the formation of the complexes Cp*M(L)Cl2 (M = Ir, L = Bmtp, 1a·Cl2; M = Rh, L = Bmtp, 1b·Cl2; M = Ir, L = Bptp, 2a·Cl2; M = Rh, L = Bptp, 2b·Cl2; M = Ir, L = Bbtp, 3a·Cl2; M = Rh, L = Bbtp, 3b·Cl2), respectively. All compounds have been characterized by elemental analysis, NMR and IR spectra. The molecular structures of Bbtp, 1a·Cl2, 1b·Cl2, 2b·Cl2 and 3b·Cl2 have been determined by X-ray crystallography.  相似文献   

12.
Assembly of 5-sulfosalicylic acid (H3L) and d10 transition metal ions (CdII, AgI) with the neutral N-donor ligands produces five new complexes: [Cd2(HL)2(4,4′-bipy)3]n·2nH2O (1), {[Cd2(μ2-HCO2)2(4,4′-bipy)2(H2O)4][Cd(HL)2(4,4′-bipy)(H2O)2]}n (2), {[Cd(4,4′-bipy)(H2O)4][HL]·H2O}n (3), [Cd(HL)(dpp)2(H2O)]n·4nH2O (4), {[Ag(4,4′-bipy)][Hhbs]}n (5) (4,4′-bipy=4,4′-bipyridine, dpp=1,3-di(pyridin-4-yl)propane, H2hbs=4-hydroxybenzenesulfonic acid, the decarboxylation product of H3L). Complex 1 adopts a 5-connected 3D bilayer topology. Complex 2 has the herring-bone and ladder chain, which are extended to a 3D network via hydrogen bonding. In 3–4 complexes, 3 is a 3D supermolecular structure formed by polymeric chains and 2D network of HL2−, while 4 gives the double-stranded chains. In 5, ladder arrays are stacked with the 2D networks of Hhbs anions in an –ABAB– sequence. Complexes 1–4 display green luminescences in solid state at room temperature, while emission spectra of 3 and 4 show obvious blue-shifts at low temperature.  相似文献   

13.
Lithiation of O-functionalized alkyl phenyl sulfides PhSCH2CH2CH2OR (R = Me, 1a; i-Pr, 1b; t-Bu, 1c; CPh3, 1d) with n-BuLi/tmeda in n-pentane resulted in the formation of α- and ortho-lithiated compounds [Li{CH(SPh)CH2CH2OR}(tmeda)] (α-2ad) and [Li{o-C6H4SCH2CH2CH2OR)(tmeda)] (o-2ad), respectively, which has been proved by subsequent reaction with n-Bu3SnCl yielding the requisite stannylated γ-OR-functionalized propyl phenyl sulfides n-Bu3SnCH(SPh)CH2CH2OR (α-3ad) and n-Bu3Sn(o-C6H4SCH2CH2CH2OR) (o-3ad). The α/ortho ratios were found to be dependent on the sterical demand of the substituent R. Stannylated alkyl phenyl sulfides α-3ac were found to react with n-BuLi/tmeda and n-BuLi yielding the pure α-lithiated compounds α-2ac and [Li{CH(SPh)CH2CH2OR}] (α-4ab), respectively, as white to yellowish powders. Single-crystal X-ray diffraction analysis of [Li{CH(SPh)CH2CH2Ot-Bu}(tmeda)] (α-2c) exhibited a distorted tetrahedral coordination of lithium having a chelating tmeda ligand and a C,O coordinated organyl ligand. Thus, α-2c is a typical organolithium inner complex.Lithiation of O-functionalized alkyl phenyl sulfones PhSO2CH2CH2CH2OR (R = Me, 5a; i-Pr, 5b; CPh3, 5c) with n-BuLi resulted in the exclusive formation of the α-lithiated products Li[CH(SO2Ph)CH2CH2OR] (6ac) that were found to react with n-Bu3SnCl yielding the requisite α-stannylated compounds n-Bu3SnCH(SO2Ph)CH2CH2OR (7ac). The identities of all lithium and tin compounds have been unambiguously proved by NMR spectroscopy (1H, 13C, 119Sn).  相似文献   

14.
Rhodium surface siloxide complexes were prepared directly by condensation of the molecular precursors ([{Rh(μ-OSiMe3)(cod)}2], [{Rh(μ-OSiMe3)(tfb)}2], [{Rh(μ-OSiMe3)(nbd)}2]) with silanol groups on silica surface (Aerosil 200 and SBA-15) and their structures were characterized by 13C and 29Si CP/MAS NMR spectroscopy. Such single-site complexes were tested for their activity in hydrosilylation of carbon–carbon double bonds with triethoxysilane, heptamethyltrisiloxane and poly(hydro,methyl)(dimethyl)siloxane. The best catalyst appeared to be cyclooctadiene ligand-containing rhodium siloxide complex immobilized on Aerosil which was recycled as many as 20 times without loss of activity and selectivity in hydrosilylation of vinylheptamethyltrisiloxane with heptamethyltrisiloxane. On the ground of CP/MAS NMR measurements it was established that the mechanism of hydrosilylation catalyzed by silica-supported rhodium siloxide complexes is different from that for the complexes in the homogeneous system.  相似文献   

15.
[TiCl2(salen)] (1) reacts with AlMe3 (1:2) to give the heterometallic Ti(III) and Ti(IV) complexes [Ti{(μ-Cl)(AlMe2)}{(μ-Cl)(AlMe2X)}(salen)] (X=Me or Cl) (2) and [TiMe{(μ-Cl)(AlCl2Me)}(salen)] (3). Addition of diethyl ether to 3 affords [Ti(Me)Cl(salen)] (4). The analogous reaction of [TiBr2(salen)] (5) gives the crystallographically characterised [Ti{(μ-Br)(AlMe2)}{(μ-Br)(AlMe2X)}(salen)] (X=Me or Br) (6) and [Ti(Me)Br(salen)] (7) in a single step, whilst the comparable reaction of [TiCl2{(3-MeO)2salen}] (8) with AlMe3 yields [Ti(Me)Cl{(3-MeO)2salen}] (9) with no evidence of titanium(III) species. Reactivity of both halide and methyl groups of 4 has been probed using magnesium reduction, SbCl5 and AgBF4 halide abstraction and SO2 insertion reactions. Hydrolysis of [Ti(Me)X(L)] complexes affords μ-oxo species [TiX(L)]2(μ-O) [X=Cl, L=salen (13); X=Br, L=salen (14); X=Cl, L=(3-MeO)2salen (15)].  相似文献   

16.
Nickel(II), palladium(II), and platinum(II) complexes of 2-(3-mesitylimidazolylidenyl)pyrimidine (L), [Ni2(μ-Cl)2(L)4][Ag2Cl4] (3), [Ni2(μ-I)2(L)4][NiI(L)2(CH3CN)]2[Ag4I8] (4), [PdCl2(L)] (5), [PdI2(L)] (6), and [PtCl(L)2][AgCl2] (7) have been obtained from the carbene transfer reactions of [Ag(L)Cl] (2). These complexes have been fully characterized by spectroscopic methods and single-crystal X-ray structure analyses. The mono(carbene)palladium and bis(carbene)platinum complexes display normal square–planar structures. Nickel complexes 3 and 4 are rare examples of paramagnetic nickel(II) complexes of N-heterocyclic carbenes having octahedral geometry.  相似文献   

17.
The stoichiometric reaction of phenylene-1,4-diaminotetra(phosphonite), p-C6H4[N{P(OC6H4OMe-o)2}2]2 (P2NФNP2) (1) with [RuCl2(p-cymene)]2 in acetonitrile produces cis,cis-[{RuCl2(CH3CN)2}2(P2NФNP2)] (2), whereas the similar reaction of 1 with [RuCl2(p-cymene)]2 in THF medium affords a tri-chloro-bridged tetrametallic complex, [{(η6-p-cymene)Ru2(μ2-Cl)3Cl}2(P2NФNP2)] (3) irrespective of the stoichiometry and reaction conditions. The formation and structure of complexes 2 and 3 are assigned through various spectroscopic and micro analysis data. The molecular structure of 2 is confirmed by single crystal X-ray diffraction study. The catalytic activities of complexes 2 and 3 have been investigated in transfer hydrogenation reactions.  相似文献   

18.
The dinuclear gem-dithiolato bridged compounds [Rh2(μ-S2Cptn)(cod)2] (1) (CptnS22− = 1,1-cyclopentanedithiolato), [Rh2(μ-S2Chxn)(cod)2] (2) (ChxnS22− = 1,1-cyclohexanedithiolato), [Rh2(μ-S2CBn2)(cod)2] (3) (Bn2CS22− = 1,3-diphenyl-2,2-dithiolatopropane) and [Rh2(μ-S2CiPr2)(cod)2] (4) (iPr2CS22− = 2,4-dimethyl-2,2-dithiolatopentane) dissolved in toluene in the presence of monodentate phosphine or phosphite P-donor ligands under carbon monoxide/hydrogen (1:1) atmosphere are efficient catalysts for the hydroformylation of oct-1-ene under mild conditions (6.8 atm of CO/H2 and 80 °C). The influence of the gem-dithiolato ligand, the P-donor co-catalyst and the P/Rh ratio on the catalytic activity and selectivity has been explored. Aldehyde selectivities higher than 95% and turnover frequencies up to 245 h−1 have been obtained using P(OMe)3 as modifying ligand. Similar activity figures have been obtained using P(OPh)3 although the selectivities are lower. Regioselectivities toward linear aldehyde are in the range 75–85%. The performance of the catalytic systems [Rh2(μ-S2CR2)(CO)2(PPh3)2]/PPh3 has been found to be comparable to the systems [Rh2(μ-S2CR2)(cod)2] at the same P/Rh ratio. The system [Rh2(μ-S2CBn2)(cod)2] (3)/P(OPh)3 has been tested in the hydroformylation-isomerization of trans-oct-2-ene. Under optimized conditions up to 54% nonanal was obtained. Spectroscopic studies under pressure (HPNMR and HPIR) evidenced the formation of hydrido mononuclear species under catalytic conditions that are most probably responsible for the observed catalytic activity.  相似文献   

19.
The reaction of {C,N-[Fe(η5-C5H5)(η5-C5H3(CH2NMe2)-2)]}Li, (FcN)Li, with zinc chloride affords the diorganozinc complex (FcN)2Zn (1). In solution, 1 appears as a mixture of rac and meso diastereomers, whereas in the solid state it crystallizes solely as a rac diastereomer. The ratio of rac/meso diastereomers in solution is solvent-, temperature- and concentration-dependent, consistent with an intermolecular exchange between diastereomers. An intramolecular dynamic phenomenon involving dissociation and recoordination of Zn---N bonds was also observed. The reaction of 1 with zinc chloride yields the monoorganozinc compound (FcN)ZnCl (2) as a slightly soluble yellow microcrystalline powder.  相似文献   

20.
Tppz (2,3,5,6-tetrakis(2-pyridyl)pyrazine) complexes [Rh(tppz)(bpy)Cl][PF6]2.acetylacetone (bpy = 2,2′-bipyridine) and [{CdCl2}2(μ-tppz)].ethylene glycol have been synthesized and characterized by elemental analyses, IR, 1H NMR, cyclic voltammetry, photoluminescence and electronic spectral studies. Solid state structures of both complexes have been determined by single-crystal X-ray crystallography. The structural determination shows that the dinuclear Cd(II) complex, [{CdCl2}2(μ-tppz)], is a 1D coordination polymer. An ORTEP drawing of [Rh(tppz)(bpy)Cl][PF6]2.acetylacetone shows that the coordination geometry around the Rh(III) center is a distorted octahedron. [{CdCl2}2(μ-tppz)] displays intraligand 1(π–π*) fluorescence and can potentially serve as a photoactive material. For the mononuclear Rh(III) complex, only a two-electron reduction process occurs at the metal with the elimination of Cl ligand. The emission of this complex is assigned as πd* phosphorescence.  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