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.
The oxidation of an anticancer drug 5-fluorouracil (5-FU) by diperiodatoargentate(III) (DPA) was carried out both in the absence and presence of osmium(VIII) catalyst in alkaline medium at 27 °C and a constant ionic strength of 0.20 mol dm−3 spectrophotometrically attached with HI-TECH SFA-12 stopped flow accessory. The oxidation products in both the cases were identified as fluoroketene and Ag(I). The stoichiometry is same in both cases, i.e., [5-FU]:[DPA] = 1:1. The reaction was of first order in both catalysed and uncatalysed cases, with respect to [DPA] and was less than unit order in [5-FU] and negative fraction in [alkali]. The order in Os(VIII) was unity. In both cases [Ag(H3IO6)2] itself is the active species of DPA. The uncatalysed reaction in alkaline medium has been shown to proceed via a DPA-5-fluorouracil complex, which decomposes in a rate determining step to give the products. In catalysed reaction, it has been shown to proceed via a Os(VIII)-5-fluorouracil complex, which further reacts with one molecule of DPA in a rate determining step to give the products. The reaction constants involved in the different steps of the mechanisms were calculated for both the reactions. The catalytic constant (kCat.const.) was also calculated for catalysed reaction at different temperatures. The activation parameters with respect to slow step of the mechanisms were computed and discussed for both the cases. The thermodynamic quantities were also determined for both reactions.  相似文献   

2.
Dichloro(4,10-dimethyl-1,4,7,10-tetraazabicyclo[5.5.2]tetradecane)chromium(III) chloride, Dichloro(4,10-dibenzyl-1,4,7,10-tetraazabicyclo[5.5.2]tetradecane) chromium(III) chloride, and Dichloro(4,11-dimethyl-1,4,8,11-tetraazabicyclo[6.6.2] hexadecane)chromium)(III) chloride have been prepared by the reaction of anhydrous chromium(III) chloride with the appropriate cross-bridged tetraazamacrocycle. Aquation of these complexes proved difficult, but Chlorohydroxo(4,11-dimethyl-1,4,8,11-tetraazabicyclo[6.6.2]hexadecane)chromium)(III) chloride was synthesized directly from chromium(II) chloride complexation followed by exposure or the reaction to air in the presence of water. The four complexes were characterized by X-ray crystal structure determination. All contain the chromium(III) ion in a distorted octahedral geometry and the macrocycle in the cis-V configuration, as dictated by the ethylene cross-bridge. Further characterization of the hydroxo complex reveals a magnetic moment of μeff = 3.95 B.M. and electronic absorbtions in acetonitrile at λmax = 583 nm (ε = 65.8 L/cm mol), 431 nm (ε = 34.8 L/cm mol) and 369 nm (ε = 17 L/cm mol).  相似文献   

3.
The kinetics of oxidation of sunset yellow (SY) by sodium-N-chloro-p-toluenesulfonamide or chloramine-T (CAT) was studied spectrophotometrically in HClO4 and NaOH media with Os(VIII) as a catalyst in the latter medium at 298 K and 303 K, respectively. In acid medium, the experimental rate law is −d[CAT]/dt = k[CAT]0[SY]0[HClO4]−0.46. Alkali accelerates the rate of reaction and the rate law takes the form −d[CAT]/dt = k[CAT]0[SY]0[NaOH]0.23[OsO4]0.84. The solvent isotope effect was studied using D2O. Benzenesulfonic acid and 1,2-naphthoquinone-6-sulfonic acid were characterized as the oxidation products of SY. Under identical set of experimental conditions in alkaline medium, Os(VIII) catalyzed reaction is about seven-fold faster than the uncatalyzed reaction. Activation parameters for the overall reaction and also with respect to catalyst have been evaluated. The observed results have been explained by plausible mechanisms and the related rate laws have been deduced.  相似文献   

4.
Titanium(II) solutions, prepared by dissolving titanium wire in triflic acid + HF, contain equimolar quantities of Ti(IV). Treatment of such solutions with excess Fe(III) or Ru(III) complexes yield Ti(IV), but reactions with Ti(II) in excess give Ti(III). Oxidations by (NH3)5Ru(III) complexes, but not by Fe(III) species, are catalyzed by titanium(IV) and by fluoride. Stoichiometry is unchanged. The observed rate law for the Ru(III)-Ti(II)-Ti(IV) reactions in fluoride media points to competing reaction paths differing by a single F, with both routes involving a Ti(II)-Ti(IV) complex which is activated by deprotonation. It is suggested that coordination of Ti(IV) to TiII(aq) minimizes the mismatch of Jahn-Teller distortions which would be expected to lower the Ti(II,III) self-exchange rate.  相似文献   

5.
[RuCl3 · nH2O] and Na(trans-[RuCl4(DMSO)2]) were reacted with 1-pyrrolidinedithiocarbamate (PDT), its S-methyl ester (PDTM), and N,N-dimethylcarbamodithioic acid methyl ester (DMDTM) in water or methanol in order to obtain the corresponding Ru(III) derivatives. Once isolated and purified, the complexes were characterized by means of elemental analysis, conductivity measurements, FT-IR and 1H NMR spectroscopy, ion electrospray mass spectrometry (ESI-MS), and thermal analyses. The crystal structure of mer-[Ru(DMDTM)(DMSO)Cl3] has been also determined by X-ray crystallography. In vitro cytotoxic activity of all the synthesized complexes was eventually evaluated on some selected human tumor cell lines.  相似文献   

6.
A series of tetramethylammonium tetrahalogenoferrates(III), [FeBr4−nCln] (n = 0, 1, 3, 4), of general formula [(CH3)4N][FeBr4−nCln], have been synthesized. The crystal and molecular structures of [(CH3)4N][FeCl4] were determined. The compound is isostructural with its [FeBr4−nCln] (n = 0, 1, 3, 4) analogues. Magnetic measurements of the powdered samples of [(CH3)4N][FeBr4−nCln] gave negative values of the Weiss constant, which suggest antiferromagnetic coupling. The strength of the antiferromagnetic interactions strongly depends on the kind of halide ligands in the coordination sphere of iron(III) and increases with an increasing number of the bromide anions.  相似文献   

7.
This work reports the synthesis, characterization, and aqueous chemistry of a series of cytotoxic [Au(polypyridyl)Cl2]PF6 complexes {(where polypyridyl = dipyrido[3,2-f:2′,3′-h] quinoxaline (DPQ), dipyrido[3,2-a:2′,3′-c] phenazine (DPPZ) and dipyrido[3,2-a:2′,3′-c](6,7,8,9-tetrahydro) phenazine (DPQC))}. The crystal structure of [Au(DPQ)Cl2]PF6 was determined as example of the series and exhibits the anticipated square planar geometry common for d8 coordination complexes. The crystals of the complex belong to the space group P21/n with a = 7.624(2) Å, b = 18.274(5) Å, c = 14.411(14) Å, β = 98.03(3)°, and Z = 4. In 1H NMR studies of these compounds in the presence of aqueous buffer, all four complexes rapidly converted to the dihydroxy species [Au(polypyridyl)(OH)2] in a stepwise fashion. However, the [Au(polypyridyl)]3+ fragment believed to impart cytotoxicity in human ovarian cancer cell lines (A2780) remained intact and appeared stable for days. It was also noted that these Au(III) complexes were readily reduced in the presence of the common biological reducing agents, reduced glutathione and sodium ascorbate. How solution and redox stability may affect the biological activity of these novel Au(III) complexes is discussed.  相似文献   

8.
Chromium exists in nuclear waste sludges and is a problematic element in the vitrification process of high-level nuclear wastes. It is therefore necessary to treat the waste sludges to remove chromium prior to vitrification, by caustic leaching or oxidation of Cr(III) to Cr(VI). The objective of this study is to investigate the effect of oligomerization of Cr(III) on its oxidation by hypochlorite in alkaline solutions.Monomeric, dimeric and trimeric Cr(III) species in solution were separated by ion exchange. The kinetics of the oxidation of the separated species by hypochlorite in alkaline solutions was studied by UV/Vis absorption spectroscopy, and compared with the oxidation by hydrogen peroxide previously studied. Results indicate that hypochlorite can oxidize Cr(III) to Cr(VI) in alkaline solutions, but the rate of oxidation by hypochlorite is slower than that by hydrogen peroxide at the same alkalinity and concentrations of oxidants. The rate of oxidation of Cr(III) by both oxidants decreases as the concentration of sodium hydroxide is increased, but the oxidation by hypochlorite seems less affected by the degree of oligomerization of Cr(III) than that by peroxide. Compared with the oxidation by hydrogen peroxide where the major reaction pathway has an inverse order with respect to CNaOH, the oxidation by hypochlorite has a significant reaction pathway independent of [OH].  相似文献   

9.
The EPR and 1H NMR spectroscopy of seven [Ru(NH3)5L]2+ complexes, where L = 3,5-dimethoxyphenylcyanamide (MeO2pcyd), 3,4,5-trimethoxyphenylcyanamide (MeO3pcyd), 4-nitrophenylcyanamide (NO2pcyd), 2,3-dichlorophenylcyanamide (Cl2pcyd), 2,4,6-trichlorophenylcyanamide (Cl3pcyd), 2,3,5,6-tetrachlorophenylcyanamide (Cl4pcyd) and pentachlorophenylcyanamide (Cl5pcyd), was performed. EPR spectra of the complexes showed an axial signal with g|| and g at high and low field, respectively. The g|| axis is suggested to lie along the Ru-cyanamide bond. Gas-phase DFT calculations of [Ru(NH3)5 phenylcyanamide]2+ showed spin density localized mostly on the phenylcyanamide ligand, in disagreement with EPR data. DFT/polarizable continuum model (PCM, water solvation) calculations shifted spin density towards ruthenium so that spin density was shared between ruthenium and phenylcyanamide ligand. Proton contact shifts were determined from NMR and EPR data and were used to estimate spin density distributions on phenyl ring carbons. The results showed that the DFT/PCM calculation overestimated spin density on phenyl ring carbons by approximately one order of magnitude. Donor-acceptor interactions between the solute and solvent that are not fully accounted for in the DFT/PCM method are suggested to stabilize the Ru(III) oxidation state.  相似文献   

10.
In this work we present the synthesis and characterization of the complex dichloro[N-propanoate-N,N-bis-(2-pyridylmethyl)amine]iron(III) [FeIII(PBMPA)Cl2]. The ligand LiPBMPA was synthesized through the Michael reaction of BMPA with methylacrylate, followed by alkaline hydrolysis. The complex [FeIII(PBMPA)Cl2] has been synthesized by the reaction of the ligand with FeCl3 · H2O and was mainly characterized by cyclic voltammetry, conductivimetry, and electronic, infrared and Mössbauer spectroscopies, and by X-ray structural analysis, which showed an iron center coordinated by one carboxylate oxygen in a monodentate way, one tertiary amine, two pyridine groups and two chloride ions. It has been proposed that in water the chloride ligands are shifted by the solvent molecules and the species [FeIII(PBMPA)(H2O)2]Cl2 is predominant. The catalase-like activity of the complex was tested in water, and it proved to be active in the hydrogen peroxide dismutation. Kinetics studies were conducted following the initial rates method. The reaction is first order in relation to both the complex and the hydrogen peroxide. Based on the presence of a lag phase that depends on the initial complex concentration, we propose that the active species that shows in situ catalase-like activity, is a binuclear complex.  相似文献   

11.
Spectroscopic studies on trans-[Ru(NH3)4(SO4)L]+ where L=imidazole, histidine, pyridine and substituted pyridines were undertaken to understand the effect of various ligands on the Ru-N bonding in these complexes. The sulfate complexes show two major bands in the 250-270 and 310-350 nm region of the UV-Vis spectrum. Based on quantum chemical calculations the lowest energy band has been assigned to a LMCT (SO4 2− → RuIII) transition. The energy of the LMCT transition decreases as the order of the axial ligand L basicity: Him > L-hist > 4-NH2-py > 4-Cl-py > 4-pic > py > nia > 4-Cn-py > isn > pz. EPR spectra give only two g values showing that the two LUMO containing the metal dπ orbitals are degenerate and the energy separation between the LUMO and HOMO, calculated from the g values correlates linearly with the charge transfer energy and electrochemical properties. These correlations suggest extensive π donation from L to the Ru(III) d orbitals. An X-ray study of the 4-pic complex shows a bent S-O-Ru bond of 127.5° and MO calculations for three other complexes predict similar angles due to extensive σ and π bonding interaction between the sulfate oxygen and the Ru(III) ion. Surprisingly, the MO calculations do not predict the observed degeneracy in the LUMO orbital found by EPR studies. We shall argue that these discrepancies can be reconciled by insisting that the orientation of the L ring be coplanar with the S-O-Ru plane as is the case in the one X-ray study.  相似文献   

12.
A heterodinuclear (Ru(II), Co(III)) metal polypyridyl complex [(phen)2Ru(bpibH2)Co(phen)2]5+ {phen = 1,10-phenanthroline, bpibH2 = 1,4-bis([1,10]phebanthroline-[5,6-d]imidazol-2-yl)-benzene} has been designed and synthesized. The comparative study on the interactions of the Ru(II)-Co(III) complex with calf thymus DNA (CT-DNA) and yeast tRNA has been investigated by UV-visible spectroscopy, fluorescence spectroscopy, viscosity, as well as equilibrium dialysis and circular dichroism (CD). The antitumor activities of the complex have been evaluated by MTT {3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide} method and Giemsa staining experiment. These results indicate that the structures of nucleic acids have significant effects on the binding behaviors of metal complexes. Furthermore, the complex demonstrates different antitumor activity against selected tumor cell lines in vitro, and can make the cell apoptosis.  相似文献   

13.
Plumbagin (5-hydroxy-2-methyl-1,4-naphthoquinone, H-PLN) was isolated from Plumbago zeylanica, the anticancer traditional Chinese medicine (TCM). Five new lanthanide(III) complexes of deprotonated plumbagin: [Y(PLN)3(H2O)2] (1), [La(PLN)3(H2O)2] (2), [Sm(PLN)3(H2O)2]⋅H2O (3), [Gd(PLN)3(H2O)2] (4), and [Dy(PLN)3(H2O)2] (5) were synthesized by the reaction of plumbagin with the corresponding lanthanide salts, in amounts equal to ligand/metal molar ratio of 3:1. The PLN-lanthanide(III) complexes were characterized by different physicochemical methods: elemental analyses, UV-visible, IR and 1H NMR and ESI-MS (electrospray ionization mass spectrum) as well as TGA (thermogravimetric analysis). The plumbagin and its lanthanide(III) complexes 1-5, were tested for their in vitro cytotoxicity against BEL7404 (liver cancer) cell lines by MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide) assay. The five PLN-lanthanide (III) complexes 1-5 effectively inhibited BEL7404 cell lines growth with IC50 values of 11.0 ± 3.5, 5.1 ± 1.3, 6.1 ± 1.1, 6.4 ± 1.3, and 9.8 ± 1.5 μM, respectively, and exhibited a significantly enhanced cytotoxicity compared to plumbagin and the corresponding lanthanide salts, suggesting a synergistic effect upon plumbagin coordination to the Ln(III) ion. The lanthanide complexes under investigation also exerted dose- and time-dependent cytotoxic activity. [La(PLN)3(H2O)2] (2) and plumbagin interact with calf thymus DNA (ct-DNA) mainly via intercalation mode, but for [La(PLN)3(H2O)2] (2), the electrostatic interaction should not be excluded; the binding affinity of [La(PLN)3(H2O)2] (2) to DNA is stronger than that of free plumbagin, which may correlate with the enhanced cytotoxicity of the PLN-lanthanide(III) complexes.  相似文献   

14.
The synthesis, characterization, and application in asymmetric catalytic cyclopropanation of Rh(III) and Ir(III) complexes containing (Sa,RC,RC)-O,O′-[1,1′-binaphthyl-2,2′-diyl]-N,N′-bis[1-phenyl-ethyl]phosphoramidite (1) are reported. The X-ray structures of the half-sandwich complexes [MCl2(C5Me5)(1P)] (M = Rh, 2a; M = Ir, 2b) show that the metal-phosphoramidite bond is significantly shorter in the Ir(III) analog. Chloride abstraction from 2a (with CF3SO3SiMe3 or with CF3SO3Me) and from 2b (with AgSbF6) gives the cationic species [MCl(C5Me5)(1,2-η-1P)]+ (M = Rh, 3a; M = Ir, 3b), which display a secondary interaction between the metal and a dangling phenethyl group (NCH(CH3)Ph) of the phosphoramidite ligand, as indicated by NMR spectroscopic studies. Complexes 3a and 3b slowly decompose in solution. In the case of 3b, the binuclear species [Ir2Cl3(C5Me5)2]+ is slowly formed, as indicated by an X-ray study. Preliminary catalytic tests showed that 3a cyclopropanates styrene with moderate yield (35%) and diastereoselectivity (70:30 trans:cis ratio) and with 32% ee (for the trans isomer).  相似文献   

15.
Four cobalt(III) polypyridyl complexes, [Co(phen)3−n(dpq)n]3+ (phen = 1,10-phenanthroline, dpq = dipyrido[3,2-f:2′,3′-h]-quinoxaline) (n = 0, 1, 2, and 3) were synthesized and the influences of the dpq ligand on the photophysical properties, electrochemical properties, DNA binding affinities, as well as photonuclease activities of the complexes, were examined in detail. The presence of dpq ligand increases the DNA binding affinities of the corresponding complexes remarkably with respect to [Co(phen)3]3+. With the sequential substitution of phen ligand by dpq ligand, the 1O2 quantum yields of the corresponding complexes are enhanced greatly. As a result, the photonuclease activities follow the order of [Co(dpq)3]3+ > [Co(phen)(dpq)2]3+ > [Co(phen)2(dpq)]3+ ? [Co(phen)3]3+. It was found all the examined complexes can generate OH upon UV irradiation, and OH is also involved in DNA photocleavage as reactive oxygen species.  相似文献   

16.
Using the 1:2 condensate of benzildihydrazone and 2-acetylpyridine as a tetradentate N donor ligand L, LaL(NO3)3 (1) and EuL(NO3)3 (2), which are pale yellow in colour, are synthesized. While single crystals of 1 could not be obtained, 2 crystallises as a monodichloromethane solvate, 2·CH2Cl2 in the space group Cc with a = 11.7099(5) Å, b = 16.4872(5) Å, c = 17.9224(6) Å and β = 104.048(4)°. From the X-ray crystal structure, 2 is found to be a rare example of monohelical complex of Eu(III). Complex 1 is diamagnetic. The magnetic moment of 2 at room temperature is 3.32 BM. Comparing the FT-IR spectra of 1 and 2, it is concluded that 1 also is a mononuclear single helix. 1H NMR reveals that both 1 and 2 are mixtures of two diastereomers. In the case of the La(III) complex (1), the diastereomeric excess is only 10% but in the Eu(III) complex 2 it is 80%. The occurrence of diastereomerism is explained by the chiralities of the helical motif and the type of pentakis chelates present in 1 and 2.  相似文献   

17.
For the first time, the Ir(III) catalysis of the iodate oxidation of xylose and maltose in aqueous alkaline medium has been investigated. The reactions exhibit first-order kinetics with respect to lower [IO(3)(-)] and [OH(-)] and show zero-order kinetics at their higher concentrations. Unity order at low concentrations of maltose becomes zero order at its higher concentrations, whereas zero-order kinetics with respect to [xylose] was observed throughout its variation. The reaction rate is found to be directly proportional to [Ir(III)] in the oxidation of both reducing sugars. Negligible effect of [Cl(-)] and nil effect of ionic strength (mu) on the rate of oxidation have also been noted. The species, [IrCl(3)(H(2)O)(2)OH](-) was ascertained as the reactive species of Ir(III) chloride for both the redox systems. Various activation parameters have been calculated. Formic acid and arabinonic acid for maltose and formic acid and threonic acid for xylose were identified as the main oxidation products of the reactions. Mechanisms consistent with the observed kinetic data and spectral evidence have been proposed for the oxidation of xylose and maltose.  相似文献   

18.
A new chemiluminescence (CL) system based on the reaction of Ag(III) complex with luminol is, for the first time, reported in this work. Incorporated with a flow injection analyses (FIA), the new CL system has been applied for the determination of free cortisol in human sera. The system is based on the CL reaction of luminol with Ag(III) in alkaline solutions, while cortisol can dramatically enhance CL intensities. Under optimum conditions, CL intensities are proportional to concentration of cortisol in the range of 0.05-7.5 nM. The limit of detection is 2.0 × 10−11 M (3σ), with a relative standard deviation (n = 11) of 1.9% for 3.5 × 10−9 M cortisol. Eight human blood serum samples were all handled by solid-phase extraction (SPE) clean-up and enrichment before detection. This detection system is highly sensitive and convenient and may find wide applications. Based on the chemiluminescent spectra, a possible reaction mechanism is also suggested.  相似文献   

19.
A study on the structure and magnetic properties of [Pr(NO3)(Pic)(H2O)2(EO3)](Pic) complex, where EO3 = triethylene glycol and Pic = picrate anion was conducted and characterized by single crystal X-ray structure analysis. Magnetic susceptibility (χM) was carried out from 2 to 300 K under both field-cooled (FC) and zero-field-cooled (ZFC) measurements with an applied magnetic field of 2000 Oe. The complex is crystallized in triclinic with space group . The coordination geometry around the Pr(III) ion was a tetradecahedron with a ten-coordination number. In the crystal, the molecular structure was stabilized by moderate and weak hydrogen bonding interactions between the cation [Pr(NO3)(Pic)(H2O)2(EO3)]+ moiety and [Pic] as counter-anion that led to the formation of a one-dimensional network. The temperature dependence of the magnetic susceptibility of [Pr(NO3)(Pic)(H2O)2(EO3)](Pic) shows the presence of weak antiferromagnetic interactions between the Pr(III) centers. The magnetic susceptibility for complex also obeys the Curie-Weiss law and is effective at high temperatures. Some factors that influence the photoluminescence intensity were also reported.  相似文献   

20.
The dissociation kinetics of the europium(III) complex with H8dotp ligand was studied by means of molecular absorption spectroscopy in UV region at ionic strength 3.0 mol dm−3 (Na,H)ClO4 and in temperature region 25-60 °C. Time-resolved laser-induced fluorescence spectroscopy (TRLIFS) was employed in order to determine the number of water molecules in the first coordination sphere of the europium(III) reaction intermediates and the final products. This technique was also utilized to deduce the composition of reaction intermediates in course of dissociation reaction simultaneously with calculation of rate constants and it demonstrates the elucidation of intimate reaction mechanism. The thermodynamic parameters for the formation of kinetic intermediate (ΔH0 = 11 ± 3 kJ mol−1, ΔS0 = 41 ± 11 J K−1 mol−1) and the activation parameters (Ea = 69 ± 8 kJ mol−1, ΔH = 67 ± 8 kJ mol−1, ΔS = −83 ± 24 J K−1 mol−1) for the rate-determining step describing the complex dissociation were determined. The mechanism of proton-assisted reaction was proposed on the basis of the experimental data.  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