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 15 毫秒
1.
The series of ferrocene-containing tris-β-diketonato aluminum(III) complexes [Al(FcCOCHCOR)(3)] (R = CF(3), 1; CH(3), 2; C(6)H(5), 3; and Fc = ferrocenyl = Fe(η(5)-C(5)H(5))(η(5)-C(5)H(4)), 4) were synthesized and investigated structurally and electrochemically; complex 1 was subjected to cytotoxicity tests. (1)H NMR-spectroscopy distinguished between the mer and fac isomers of 2 and 3. Complex 1 existed only as the mer isomer. A single crystal X-ray crystallographic determination of the structure of a mer-isomer of Al(FcCOCHCOCF(3))(3), 1, (Z = 4, space group P2(1)2(1)2(1)) demonstrated extensive delocalization of all bonds which explained the pronounced electrochemically observed intramolecular communication between molecular fragments. In contrast to electrochemical studies in CH(2)Cl(2)/[N((n)Bu)(4)][PF(6)], the use of the supporting electrolyte [N((n)Bu)(4)][B(C(6)F(5))(4)] allowed identification of all Fc/Fc(+) electrochemical couples by cyclic and square wave voltammetry for 1-4. For R = Fc, formal reduction potentials of the six ferrocenyl groups were found to be E°' = 33, 123, 304, 432, 583, and 741 mV versus free ferrocene respectively. Complex 1 (IC(50) = 10.6 μmol dm(-3)) was less cytotoxic than the free FcCOCH(2)COCF(3) ligand having IC(50) = 6.8 μmol dm(-3) and approximately 2 orders of magnitude less toxic to human HeLa neoplastic cells than cisplatin (IC(50) = 0.19 μmol dm(-3)).  相似文献   

2.
Treatment of MeOH solutions of [Rh(cod)(fca)] (cod = 1,5-cyclooctadiene, fca = ferrocenoylacetonato) with seven derivatives of 1,10-phenanthroline (N,N), as well as with the (N,N) ligand 2,2-dipyridyl, gave [Rh(cod)(N,N)]+. The kinetics of these reactions follow the rate law: Rate = k[Rh(cod)(fca)[N,N] The temperature dependence of all the studied substitutions resulted in activation entropies, S , more negative than –100 J K–1 mol–1 which is indicative of associative mechanisms. The pK a's of the incoming phenanthroline derivatives were between 3.03 and 6.31 but did not influence the reaction rate to any significant extent. This implies that the rate determining step during the substitution involves Rh—O bond breaking and not Rh—N bond formation. Substitution of fca with 2,2-dipyridyl was slightly faster (k = 118 dm3 mol–1 s–1) than with the 1,10-phenanthroline derivatives (k average = 14.2 dm3 mol–1 s–1) and may be attributed to the free rotation capability of the two pyridyl rings about the 1-1 carbon–carbon axis in 2,2-dipyridyl. 1,10-Phenanthroline cannot rotate about the corresponding carbon axis.  相似文献   

3.
The syntheses of the novel cyclopentadienylphosphinevinylidenerhodium complexes C5H5Rh(CCHR)(PPri3) (R = Ph, Me, H) and, for R = Ph, of the isomeric alkynyl hydrido compound C5H5 RhH(C2Ph)(PPri3) are reported. The square-planar complexes trans-[RhCl(RC2H)(PPri3)2] (IIa, IIb), which in solution are in equi-librium with the five-coordinated pyridine to give the octahedral compounds RhHCl(C2R)(PRri3)2(py) (VIa, VIb). Treatment of Via, VIb with NaC5H5 gives the vinylidene complexes IVa, IVb in good yield. C5H5Rh(CCH2)(PPri3) (IVc) is directly obtained from trans-[RhCl(C2H2)(PPri3)2] (IIc) and NaC5H5. Mechanistic studies confirm that the reaction of VIa, VIb with the cyclopentadienide anion primarily gives, by elimination of HCl, the rhodium(I) compounds trans-[Rh(C2R)(py)(PPri3)2] (VIIIa, VIIIb), which react with cyclopentadiene, possibly via trans-[Rh(C2R)(η2-C5H6)(PPri3)2](X) as an intermediate, to give C5 VIIIa with cyclopentadiene in presence of water gives the complex C5H5RhH(C2Ph)(PPri3), which isomerizes only slowly to form IVa and, therefore, is not an intermediate in the reaction of VIIIa and C5H6 to give IVa. The crystal structure of IVa has been determined. The RhCC arrangement is almost linear. The RhC distance is significantly shorter than in carbenerhodium complexes, which, in agreement with 13C NMR data and MO calculations, indicates a high degree of multiple bonding.  相似文献   

4.
The 2-picolylcyclopentadienyl derivatives of rhodium(I) and iridium(I) of formula [M{η5-C5H4(2-CH2C5H4N)}(η4-C8H12)] (3) (M = Rh) and (4) (M = Ir) are obtained in good yields by reacting 2-picolylcyclopentadienyllithium (7) with [RhCl(η4-C8H12)]2 and [IrCl(η4-C8H12)]2, respectively. The corresponding dicarbonyl derivatives, [M{η5-C5H4(2-CH2C5H4N)}(CO)2] (5) (M = Rh) and 6 (M = Ir), are obtained in good yields by reacting 2-picolylcyclopentadienylthallium(I) (8) with [RhCl(CO)2]2 and [IrCl(C5H5N)(CO)2], respectively. 5 has already been reported in the literature. The new complexes were characterized by elemental analysis, mass spectrometry, 1H NMR, FT-IR, and UV-Vis (210-330 nm) spectroscopy. The UV-Vis spectra indicate the existence of some electronic interaction between the 2-picolinic chromophore and the cyclopentadienyl-metal moiety. The study of the electrochemical behaviour of 3-6 by cyclic voltammetry (CV) allows the interpretation of the electrode processes and gives information about the location of the redox sites. Moreover, various synthetic strategies were tested in order to try to coordinate the complexes 3-6 to a ruthenium(II) centre, but most of them failed. Instead, the hetero-bimetallic complex bis(2,2′-bipyridine)[(η5-2-picolylcyclopentadienyl)(η4-cycloocta-1,5-diene)rhodium(I)]chlororuthenium(II)-(hexafluorophosphate) (13), was obtained, although in poor yields (10%), by reacting the nitrosyl complex [RuCl(bipy)2(NO)][PF6]214 (bipy = 2,2′-bipyridine) first with potassium azide and then with the rhodium(I) complex 3. The analogous complex bis(2,2′-bipyridine)(2-picoline)chlororuthenium(II)-(hexafluorophosphate) (15), that carries a ruthenium-bonded 2-picoline molecule instead of 3, has prepared in the same way. 13 and 15 were characterized by elemental analysis, mass spectrometry, and 1H NMR.  相似文献   

5.
The ease of formation of the phosphonate complexes [M(P(O)(OMe)2(P(OMe)3)4] (M = Rh, Ir), from the pentakis-trimethylphosphite complexes [M(P(OMe)3)5]Cl is reported. Differences in the interaction of H2 with the complexes [M′(P(O)(OMe)2(P(OMe)3)4), (M′ = Co, Rh, Ir) are presented and discussed.  相似文献   

6.
Low temperatùre (−70°C) reaction of (cod)2Rh22 (μ-Cl)2 with two molar equivalents of RLi in diethyl ether gives (cod)2Rh2 (μ-R)2 where R = Me, Me3SiCH2. Even though the bridging alkyls are air and moisture sensitive, they may be stored for prolonged periods at −30°C. The bridging alkyls are fluxional at + 20°C and the NMR spectra are consistent with a dimer of idealized D2h symmetry. Low temperature NMR spectroscopic studies suggest that the dimers have idealized C2v symmetry as found by X-ray studies described earlier. The bridging alkyls readily react with Lewis bases to give monomeric (cod)Rh(R)(L) where R = Me and L = PMe3, PEt3, P(NMe2)3, P(OMe)3, py and R = Me3SiCH2, L = P(OMe)3. The five coordinate, fluxional complex, (cod)RhMe(PMe3)2 also may be isolated. The four coordinate (cod)RhMe(PEt3), slowly reacts with toluene to give (cod)Rh(m-tolyl)(PEt3), (cod)Rh(p-tolyl)(PEt3), and methane and (cod)RhMe(PMe3) slowly reacts with benzene to give (cod)Rh(Ph)(PMe3) and methane.  相似文献   

7.
《Tetrahedron: Asymmetry》1999,10(15):3039-3043
Complexes of (R)-BINAP (BINAP=2,2′-bis(diphenylphosphino)-1,1′-binaphthyl) derived from the available rhodium precursors Rh(acac)(CO)2 and [Rh(μ-OMe)(cod)]2 are used for the asymmetric hydroformylation of vinyl acetate. Enantiomeric excesses of up to 60% are achieved with regioselectivities of up to 99%. Only a BINAP/Rh ratio of 2 is required. Effects of pressure and temperature on catalyst stability, enantio- and chemoselectivity are discussed.  相似文献   

8.
Summary The thiolato-bridged dinuclear compounds [Rh(-SR)-(COD)]2, where R=p-C6HF4 (1),p-C6H4F (2) and CF3 (3), are obtained from the chloro-bridged analogue by ligand exchange.Compound (1) crystallizes in the space group P1 with a=9.740(3)Å, b=11.642(4)Å, c=13.997(6)Å, =103.87(3)°, =106.98(3)° and =105.10(2)°; z=2. In this dinuclear molecule both Rh atoms have a square planar coordination sharing one edge, namely the two sulphur bridging atoms. The Rh—Rh separation of 2.96 Å is consistent with at most a very weak metal-metal interaction. Upon addition of CO the dimeric [Rh(-SR)(CO)2]2 (4), (5) and (6) are obtained, but addition of PPh3 affords the monomeric species [Rh(SR)(PPh3)-(COD)] (7), (8) and (9). Reactions of the dimeric tetracarbonyl derivatives with PPh3 vary with the nature of R; [Rh(-SR)(PPh3)(CO)]2 is obtained when R=p-C6H4F (10) and CF3 (11) but monomeric [Rh(SR)-(PPh3)(CO)2] (12) is produced when R=p-C6HF4. The latter mononuclear compounds, with R=p-C6H4F (13) and CF3 (14), are also formed by reaction of [Rh(SR)-(PPh3)(COD)] with CO.  相似文献   

9.
New rhodium(I)- and rhodium(III)-β-diketonato complexes of the type [Rh(FcCOCHCOR)(P(OPh)3)2] and [Rh(FcCOCHCOR)(P(OPh)3)2(CH3)(I)], with Fc = ferrocenyl and R = Fc, CH3 and CF3, have been synthesized. The reactivity of complexes of the type [Rh(β-diketonato)(P(OPh)3)2] increase in the order: β-diketonato = (CF3COCHCOCF3) < (CF3COCHCOPh) < (CF3COCHCOCH3) < (PhCOCHCOPh) < (CF3COCHCOFc) < (CH3COCHCOPh) < (CH3COCHCOCH3) < (CH3COCHCOFc) < (FcCOCHCOFc), giving linear relationships between the kinetic parameter ln k2 and the parameters that are related to the electron density on the rhodium centre; the sum of the group electronegativities of the β-diketonato side groups (χR + χR′) and the pKa of the uncoordinated β-diketone RCOCH2COR′. The large negative values of the volume and entropy of activation indicated a mechanism which occurs via a polar transition state. A density functional theory study, at the PW91/TZP level of theory, indicates that oxidative addition of iodo methane to [Rh(FcCOCHCOCF3)(P(OCH3)3)2] occurs via a two-step mechanism. This mechanism involves a nucleophilic attack by the metal on the methyl carbon to displace iodide to form a metal-carbon bond and the coordination of iodide to the five-coordinated intermediate to give a six-coordinated trans alkyl product.  相似文献   

10.
The effect of the nature of β-diketones on the spectrochemical-luminescence properties of lanthanide complexes used as analytical forms for the highly sensitive determination of some rare-earth elements was examined. NMR spectroscopy, molecular mechanics, calculations, and an experimental evaluation of the hydrophobicity of the test β-diketones indicated that an increase in the intensity, quantum yield, and lifetime of luminescence of lanthanide (Ln) β-diketonates depends on the charge distribution in the chelate ring, on the spatial structure, and on the hydrophobicity of the coordinated ligand. These factors determine the efficiency of energy transfer from the β-diketone to the Ln+3 ion and the decrease in energy losses caused by the quenching action of water molecules.  相似文献   

11.
Lead(II) 4,4,4-trifluoro-1-phenyl-1,3-butandionate (TFPB?) complexes with 1,10-phenanthroline (phen) and 2,2′-bipyridine (2,2′-bipy), [Pb(L)(TFPB)2], have been synthesized and characterized by elemental analysis, IR-, 1H NMR spectroscopy and studied by X-ray crystallography. The self-assembly of [Pb(L)(TFPB)2] complexes, (L?=?phen or 2,2′-bipy) is caused by CH?···?F–C, C–H?···?O–C and π–π stacking interactions. The thermal stabilities of compounds were studied by thermal gravimetric (TG) and differential thermal analyses (DTA).  相似文献   

12.
13.
Volatile iridium(I) complexes [Ir(cod)Cpx] (Cpx = pentamethylcyclopentadienyl Cp*, ethylcyclopentadienyl CpEt, cod = 1,5-cyclooctadiene) are synthesized and characterized by IR and NMR spectroscopy. The [Ir(cod)Cp*] complex is a solid and the [Ir(cod)CpEt] complex is a liquid (SATP). The XRD method is used to determine the structure of the [Ir(cod)Cp*] complex: chemical formula C18H27Ir, space group P21/c, a = 8,4418(2) Å, b = 9,4764(3) Å, c = 19.2682(5) Å, β = 96.128(1) °, V = 1532.61(7) Å3, Z = 4, d calc = 1.888 g/cm3, μ = 8.697 mm–1. The cyclopentadienyl ligand is η5-type coordinated; 1,5-cyclooctadiene have a cis-cis conformation and is η4-type coordinated. The thermal properties of the complexes are studied by thermogravimetry.  相似文献   

14.
《Polyhedron》1999,18(26):3425-3431
Copper(I) complexes with di-2-pyridylketone oxime (DPKox) of the type CuLX·nH2O, n=1 for X=Cl and Br, and n=0 for X=I and SCN, have been synthesized and characterized. The overall physical results suggest tridentate and bidentate DPKox ligand in the Cl, Br and I, SCN complexes, respectively, and terminal X in the former but bridging X in the later. These complexes display MLCT bands in the visible region, but they do not fluoresce at room temperature. The structure determination has shown the chloride complex (1) to have a centro-symmetrically related dimeric unit, in which each copper atom is coordinated by Cl(1), N(1), N(2) and N(3) (of the second ligand molecule) in a distorted tetrahedral environment. Hydrogen bonds are formed by the O(1) of the oxime group and a lattice water molecule, and between different lattice water molecules and Cl(1). The structure of the thiocyanate complex (2) features tetrahedral geometry around copper atoms, a chelating bidentate DPKox ligand coordinating via one of the two pyridyl nitrogens, N(1), and N(oxime) only and μ-1,3-thiocyanate group forming zigzag chains along the c-axis of the unit cell.  相似文献   

15.
《Comptes Rendus Chimie》2002,5(5):431-440
Several novel rhodium allyl complexes have been prepared and their structures have been studied using NMR spectroscopy and X-ray crystallography. Depending on the bite angle of the ligand and the substitution pattern of the allyl group, two different coordination modes (η1 and η3) have been observed for the allyl moiety. The activities of these Rh allyl complexes in the allylic alkylation reaction have been tested. We have shown that both coordination modes give active complexes in this reaction, but that the regioselectivity is dependent on the coordination mode of the allyl group.  相似文献   

16.
《Polyhedron》1986,5(3):633-641
An equilibrium study has been carried out on the interaction of ethionine(eth) with Pd(II) in aqueous solution at I = 0.16 M (Cl and 25°C using potentiometic methods. It has been concluded that five complex species exist in the pH range 2.8–4.8. these species are: PdCl3(Eth0H02, PdCl2(Eth), PdClOH(Eth), Pd(Eth)2(H)2+2 and Pd(Eth)02. In addition, the stopped-flow method has been used to study the reaction kinetics of Pd(II) with Eth. Three kinetic steps were observed in the pH range 1–5.5. These steps are dependent on the total concentration of Eth (TEth) as well as the pH of the medium. The observed pseudo-first order rate constants for the three reaction kinetic steps at constant pH are expressed empirically by kiobs = mi + miTEth. The parameters mi and mi are pH-dependent. It has been concluded that PdCl2−4 and PdCl2OH2− species play an important role in the complex formation reactions with Eth. The data were interpreted in terms of the complex species obtained from the equilibrium study. cis-trans substitution reactions have been suggested to account for some kinetic steps.  相似文献   

17.
Experimental redox potentials of 16 derivatives of tris(β-diketonato)iron(III) complexes (where β-diketonato(R1COCHCOR2), with substituents R1 and R2 in different combinations of H, C4H3S, C4H3O, CH3, Ph, CF3, or C (CH3)3), and 11 additional isomers, were studied theoretically in terms of the electronic properties, substituent effects, electron affinity, and molecular electrostatic potential (MESP) analysis, using density functional theory methods. The computational methods reproduced the experimental reduction potential to a very high level of accuracy, especially when the M062X functional was used (with mean absolute deviation [MAD] = 0.054 and 0.093 and correlation R2 = 0.978 and 0.981 obtained by application of two slightly different free energy cycles, respectively). The most negative computed reduction potential corresponds to the most negative reported experimental reductions, which is indicative of the least favorable reduction potential, also in most cases the most stable molecules energetically. The calculated reduction potentials of the fac isomers of the molecules were generally higher (less negative) than that of the mer isomers when one of the ligand substituents R1 or R2 was CF3 (M062X results), indicating better ease of reduction, although in many cases, the experimental reduction potential agreed better with the calculated reduction potential of the mer isomer instead. The calculated reduction potentials were also affected by the substituents in the order of CF3 > H > C4H3S > C4H3O > Ph > CH3 > C(CH3)3 (the most negative value). The stronger the electron withdrawing tendency of the substituent, the more favorable (less negative value) the reduction potential becomes. In relation to the CH3-substituted molecule 1 as a reference, the molecules with electron withdrawing substituents resulted in an electron-deficient MESP iso-surface, in both the neutral state and reduced state. All the molecules in their reduced state were characterized with an electron-deficient MESP iso-surface compared with the reduced CH3-substituted molecule 1, with the deficiency increasing in mer compared with fac, for both the neutral and reduced molecules. The relative MESP values of ΔVFe in the reduced state of the molecules were able to predict the corresponding experimental reduction potential to a significant level of accuracy (with MAD = 0.091).  相似文献   

18.
19.
In this study, new rhodium(I) complexes (5 and 6) have been prepared by the reaction of [RhCl(COD)]2 with a series of imidazolium salts (3 and 4), which were obtained from a chiral amino alcohol. The catalytic activities of these complexes were tested in the arylation of aldehydes. It was found that the synthesized rhodium complexes were highly effective catalysts for the arylation of aldehydes in short reaction times (5 min, TOF = 1193 h−1). However, the obtained ee values (up to 32% ee) remained low. We have proposed a mechanism for the arylation reaction of aldehydes, which is confirmed via 19F NMR spectroscopy.  相似文献   

20.
The substitution of the CO ligand in rhodium(I) β-ketoiminato complexes Rh(R1{O,N}R2)(CO)2 ({O,N}=R1C(O)CHC(NH)R2; R1, R2=CF3, Me, CMe3 in several combinations) by phosphorus ligands PZ3 (PZ3=PCy3, PPh3, P(OPh)3, P(NC4H4)3) leads to Rh(R1{O,N}R2)(CO)(PZ3) complexes characterised by 31P{1H}-NMR and X-ray methods. The stronger σ-donor PZ3 ligands (PZ3=PCy3, PPh3) substitute almost exclusively the CO group trans to N, forming P-trans-to-N isomers. The complexes Rh(CF3{O,N}Me)(CO)(PCy3) (II), Rh(CF3{O,N}CMe3)(CO)(PCy3) (III), Rh(CF3{O,N}Me)(CO)(PPh3) (IV) and Rh(CF3{O,N}CMe)(CO)(PPh3) (V) are of a square-planar geometry with a slight tetrahedral distortion around the rhodium atom in II, III and V. The RhP(PCy3) bonds are slightly longer than the RhP(PPh3) bonds. The reaction of stoichiometric amounts of the less basic P(OPh)3 or P(NC4H4)3 ligands leads to the formation of both isomers of the Rh(R1{O,N}R2)(CO)(P(OPh)3) or Rh(R1{O,N}R2)(CO)(P(NC4H4)3) complex in comparable yields. The RhP(P(OPh)3) distance (2.195(2) Å) in the isomer of Rh(CF3{O,N}CMe3)(CO)(P(OPh)3) with P(OPh)3 coordinated trans to N (VI) is ca. 0.04 Å longer than in the isomer of that complex with P(OPh)3 coordinated trans to O (VII). The CO substitution in Rh(R1{O,N}R2)(CO)2 by PZ3 ligands (PPh3, PCy3, P(OPh)3) causes the shortening of the RhC(CO) bond by ca. 0.04 Å compared to Rh(CF3{O,N}Me)(CO)2 (I), making difficult the coordination of another PZ3 ligand, especially one with stronger σ-donor properties. The more π-acceptor P(OPh)3 ligands form bis-phosphito complexes and Rh(CF3{O,N}CMe3){P(OPh)3}2 (VIII) exhibits inequivalence of the two P(OPh)3 ligands in solution (31P-NMR) as well as in solid form (X-ray).  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