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 281 毫秒
1.
A systematic study of the reduction of (ImH)[trans-RuCl4(dmso)(Im)] (NAMI-A; dmso is dimethyl sulfoxide, Im is imidazole), a promising antimetastasing agent entering phase II clinical trial, by l-ascorbic acid is reported. The rapid reduction of trans-[RuIIICl4(dmso)(Im)] results in formation of trans-[RuIICl4(dmso)(Im)]2− in acidic medium (pH = 5.0) and is followed by successive dissociation of the chloride ligands, which cannot be suppressed even in the presence of a large excess of chloride ions. The reduction of NAMI-A strongly depends on pH and is accelerated on increasing the pH. Over the small pH range 4.9−5.1, the reaction is quite pH-independent and the influence of temperature and pressure on the reaction could be studied. On the basis of the reported activation parameters and other experimental data, it is suggested that the redox process follows an outer-sphere electron transfer mechanism. A small contribution from a parallel reaction ascribed to inner-sphere reduction of aqua derivatives of NAMI-A, was found to be favored by lower concentrations of the NAMI-A complex and higher temperature. In the absence of an excess of chloride ions, the reduction process is catalyzed by the Ru(II) products being formed. The reduction of NAMI-A is also catalyzed by Cu(II) ions and the apparent catalytic rate constant was found to be 1.5 × 106 M−2 s−1 at 25 °C. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

2.
The coordination of the antimetastatic agent NAMI-A, [H(2)im][trans-RuCl(4)(dmso-S)(Him)], (Him=imidazole; dmso=dimethyl sulfoxide), to the DNA model base 9-methyladenine (9-MeAde) was investigated in water. NMR spectroscopy was first applied for the study of the molecular stability and hydrolysis of NAMI-A in aqueous solution over a range of pH (3.0-7.4) and chloride ion concentrations (0-1 M) at 37.0 degrees C. In physiological conditions (phosphate buffer, pH 7.4) NAMI-A disappears from the solution in 15 min due to chloride and dmso hydrolysis, leading to uncharacterised poly-oxo Ru species. Conversely, at lower pH (3.0-6.0) and in water (pH approximately 5.5), only a partial dmso hydrolysis occurs, slowly forming the [trans-RuCl(4)(H(2)O)(Him)](-) complex. This latter species coordinates to 9-MeAde (via the N7 of 9-MeAde), forming the [trans-RuCl(4)(9-MeAde)(Him)](-) complex. NAMI-A and [trans-RuCl(4)(H(2)O)(Him)](-) give comparable intracellular ruthenium concentrations and accumulate in KB cells (human mouth carcinoma) and accumulate these at the G(2)/M phase, while poly-oxo Ru species do not, and their cell uptake is reduced to 50%. On the contrary, G(2)/M arrest and protein content in the murine metastatic cell line metGM, are not influenced by NAMI-A hydrolysis. Hydrolysed NAMI-A species apparently are easier taken up by the metGM cells, showing intracellular ruthenium concentrations one order of magnitude greater than those of intact NAMI-A. Therefore, it is proposed that the selective antimetastatic activity of NAMI-A during in vivo experiments can be attributed to its hydrolysed species.  相似文献   

3.
[ImH][trans-RuIIICl4(DMSO)(Im)] (where DMSO is dimethyl sulfoxide and Im is imidazole) (NAMI-A) is an antimetastatic prodrug currently in phase II clinical trials. The mechanisms of action of this and related Ru-based anticancer agents are not well understood, but several cellular targets have been suggested. Although Ru has been observed to bind to DNA following in vitro NAMI-A exposure, little is known about Ru–DNA interactions in vivo and even less is known about how this or related metallodrugs might influence cellular RNA. In this study, Ru accumulation in cellular RNA was measured following treatment of Saccharomyces cerevisiae with NAMI-A. Drug-dependent growth and cell viability indicate relatively high tolerance, with approximately 40% cell death occurring at 6 h for 450 μM NAMI-A. Significant dose-dependent accumulation of Ru in cellular RNA was observed by inductively coupled plasma mass spectrometry measurements on RNA extracted from yeast treated with NAMI-A. In vitro, binding of Ru species to drug-treated model DNA and RNA oligonucleotides at pH 6.0 and 7.4 was characterized by matrix-assisted laser desorption/ionization time-of-flight mass spectrometry in the presence and absence of the reductant ascorbate. The extent of Ru–nucleotide interactions increases slightly with lower pH and significantly in the presence of ascorbate, with differences in observed species distribution. Taken together, these studies demonstrate the accumulation of aquated and reduced derivatives of NAMI-A on RNA in vitro and in cellulo, and enhanced binding with nucleic acid targets in a tumorlike acidic, reducing environment. To our knowledge, this is also the first study to characterize NAMI-A treatment of S. cerevisiae, a genetically tractable model organism.  相似文献   

4.
Biomimetic pathways for the oxidation of [Au(CN)(2)](-), a gold metabolite, and further cyanation of the gold(III) products to form Au(CN)(4)(-) were investigated using 13C NMR and UV-Visible spectroscopic methods. Hypochlorite ion, an oxidant released during the oxidative burst of immune cells, was employed. The reaction generates mixed dicyanoaurate(III) complexes, trans-[Au(CN)(2)X(2)](-), where X(-) represents equilibrating hydroxide and chloride ligands, and establishes the chemical feasibility of dicyanoaurate oxidation by OCl(-) to gold(III) species. This oxidation reaction suggests a new procedure for synthesis of H[Au(CN)(2)Cl(2)]. Reaction of trans-[Au(CN)(2)X(2)](-) (X(-)=Cl(-) and Br(-)) or [AuCl(4)](-) with HCN in aqueous solution at pH 7.4 leads directly to [Au(CN)(4)](-) without detection of the anticipated [Au(CN)(x)X(4-x)](-)intermediates, which is attributed to the cis- and trans-accelerating effects of the cyanides. The reduction of [Au(CN)(4)](-) by glutathione and other thiols is a complex, pH-dependent process that proceeds through two intermediates and ultimately generates [Au(CN)(2)](-). These studies provide further insight into the possible mechanisms of an immunogenically generated gold(I)/gold(III) redox cycle in vivo.  相似文献   

5.
Irradiation of trans-[RuCl(cyclam)(NO)](2+), cyclam is 1,4,8,11-tetraazacyclotetradecane, at pHs 1-7.4, with near UV light results in the release of NO and formation of trans-[Ru(III)Cl(OH)(cyclam)](+) with pH dependent quantum yields (from approximately 0.01 to 0.16 mol Einstein(-1)) lower than that for trans-[RuCl([15]aneN(4))(NO)](2+), [15]aneN(4) is 1,4,8,12-tetaazacyclopentadecane, (0.61 mol Einstein(-1)). After irradiation with 355 nm light, the trans-[RuCl([15]aneN(4))(NO)](2+) induces relaxation of the aortic ring, whereas the trans-[RuCl(cyclam)(NO)](2+) complex does not. The relaxation observed with trans-[RuCl([15]aneN(4))(NO)](2+) is consistent with a larger quantum yield of release of NO from this complex.  相似文献   

6.
Interest in Ru anticancer drugs has been growing rapidly since NAMI-A ((ImH(+))[Ru(III)Cl(4)(Im)(S-dmso)], where Im = imidazole and S-dmso = S-bound dimethylsulfoxide) or KP1019 ((IndH(+))[Ru(III)Cl(4)(Ind)(2)], where Ind = indazole) have successfully completed phase I clinical trials and an array of other Ru complexes have shown promise for future development. Herein, the recent literature is reviewed critically to ascertain likely mechanisms of action of Ru-based anticancer drugs, with the emphasis on their reactions with biological media. The most likely interactions of Ru complexes are with: (i) albumin and transferrin in blood plasma, the former serving as a Ru depot, and the latter possibly providing active transport of Ru into cells; (ii) collagens of the extracellular matrix and actins on the cell surface, which are likely to be involved in the specific anti-metastatic action of Ru complexes; (iii) regulatory enzymes within the cell membrane and/or in the cytoplasm; and (iv) DNA in the cell nucleus. Some types of Ru complexes can also promote the intracellular formation of free radical species, either through irradiation (photodynamic therapy), or through reactions with cellular reductants. The metabolic pathways involve competition among reduction, aquation, and hydrolysis in the extracellular medium; binding to transport proteins, the extracellular matrix, and cell-surface biomolecules; and diffusion into cells; with the extent to which individual drugs participate in various steps along these pathways being crucial factors in determining whether they are mainly anti-metastatic or cytotoxic. This diversity of modes of action of Ru anticancer drugs is also likely to enhance their anticancer activities and to reduce the potential for them to develop tumour resistance. New approaches to metabolic studies, such as X-ray absorption spectroscopy and X-ray fluorescence microscopy, are required to provide further mechanistic insights, which could lead to the rational design of improved Ru anticancer drugs.  相似文献   

7.
The reaction between [PtCl(dmso)(en)]Cl (dmso=dimethyl sulfoxide, en=ethylenediamine) and N-(3-pyridyl)-2-(4-(trifluoromethyl)phenyl)diazenecarboxamide (L) was studied using multinuclear NMR spectroscopy. The water-soluble complexes [PtCl(en)(L-N1)](+) (1) and [Pt(en)(L-N1)(2)](2+) (2) were isolated and their reactions with glutathione (GSH) were investigated to assess the oxidation properties of coordinated L. Both species 1 and 2 oxidized GSH to GSSG, while the reduced form of L (semicarbazide, SL) remained coordinated to Pt(2+). In complex 1 the labile chloride ion was substituted by the thiol moiety of GSH, which gave rise to the release of en in excess GSH over a period of 7 days. Complexes [PtCl(dmso)(en)]Cl, 1, 2 and ligand L were tested against T24 bladder carcinoma cells. Ligand L and complexes 1 and 2 showed higher cytotoxicity than [PtCl(dmso)(en)]Cl.  相似文献   

8.
A novel class of dianionic Ru(III) dimers of formula Na2[[trans-RuCl4(Me2SO)]2(mu-L)], with L = pyrazine (pyz, 1), pyrimidine (pym, 2), 4,4'-bipyridine (bipy, 3), and 1,2-bis(4-pyridine) ethane (etbipy, 4), was developed by us with the specific aim of assessing their antitumor properties. The dimers are in fact structurally related to the antimetastatic mononuclear compound (ImH) [trans-RuCl4(Me2SO)(Im)] (NAMI-A, Im = imidazole). Preliminary results concerning the antineoplastic activity of 1-4 against the murine MCa carcinoma model, a tumor which spontaneously metastasizes in the lungs, are reported. Similarly to what is normally observed with NAMI-A, the treatment with the dimeric complexes was scarcely effective against the growth of the primary tumor. However, dimers 1, 2, and 4 reduced very effectively the number and, in particular, the weight of lung metastases (to about 5% with respect to controls); in particular, Na2[[trans-RuCl4(Me2SO)]2(mu-etbipy)] (4) was as effective as NAMI-A in reducing the spontaneous metastases at a dosage which, in terms of moles of ruthenium, is about 3.5 times lower compared to that normally used for NAMI-A. Furthermore, in vitro tests showed that dimers 1-4 are capable of forming interstrand cross-links with linearized plasmidic DNA in a time-dependent manner. All the dimeric species are more active in inducing cross-links compared to NAMI-A, and the dimer bridged by the etbipy ligand (4) is the most effective among those tested.  相似文献   

9.
The reaction of trans-[Ru(NH(3))(4)P(OEt)(3)NO](3+) and mitochondria was investigated through differential pulse polarography and fluorimetry. The nitrosyl complex undergoes one-electron reduction centered on the NO ligand site. The reaction between the mitochondrial reductor and trans-[Ru(NH(3))(4)P(OEt)(3)NO](3+) exhibits a second order specific rate constant calculated as k=2 x 10(1) M(-1) s(-1). The reduced species, trans-[Ru(NH(3))(4)P(OEt)(3)NO](2+), quickly releases NO, yielding trans-[Ru(NH(3))(4)P(OEt)(3)H(2)O](2+). The low toxicities of both trans-[Ru(NH(3))(4)P(OEt)(3)(NO)](2+) and trans-[Ru(NH(3))(4)P(OEt)(3)H(2)O](2+) and its ability to release NO after reductive activation in a biological medium make the nitrosyl compound a useful model of a hypotensive drug.  相似文献   

10.
An electrochemical and biological study of interaction between the prototypical antimetastatic drug imidazolium trans-tetrachlorodimethylsulfoxideimidazoleruthenate (III) complex, [ImH][RuCl(4)(DMSO)(Im)] (DMSO = dimethylsulfoxide, Im = imidazole), nicknamed NAMI-A, and several biomolecules, namely DNA, bovine (BSA) and human (HSA) serum albumin, is reported. Electrochemistry offers great advantages over the existing devices based on optical techniques, since it provides rapid, simple, and low-cost information whether the interaction occurs or not. Moreover, we describe some biochemical assays to test the interaction of NAMI-A with ribonucleoprotein telomerase and protein Taq polymerase. All the data confirm the preferential interaction of NAMI-A with proteins with respect to nucleotides, especially when compared with the behaviour of the well-known alkylating drug cisplatin in the presence of the same targets.  相似文献   

11.
12.
Two new Ru(II) complexes [Ru(L)(4)(dppz)](2+) (L=imidazole (Im), 1-methylimidazole (MeIm); dppz=dipyrido[3,2-a:2',3'-c]phenazine), have been synthesized and characterized in detail by elemental analysis, (1)H NMR, Electrospray ionization mass spectrometry (ESI-MS) and UV-visible (UV-Vis) spectroscopic techniques. The interaction of these complexes with calf thymus DNA (CT-DNA) has been explored by using electronic absorption titration, competitive binding experiment, circular dichroism (CD), thermal denaturation and viscosity measurements. The experimental results show that: both the two complexes can bind to DNA in an intercalation mode; the DNA-binding affinity of complex [Ru(Im)(4)(dppz)](2+)1 (K(b)=2.5 x 10(6)M(-1)) is greater than that of complex [Ru(MeIm)(4)(dppz)](2+)2 (K(b)=1.1 x 10(6)M(-1)). Moreover, it is very interesting to find that the circular dichroic spectrum of DNA-complex 1 adduct, in which both bands centered at 277 nm and 236 nm are all negative, is very different from those of DNA-complex 2 adduct and other Ru(II) complexes binding to DNA in general intercalation mode. It may be due to the hydrogen-bonding effect or the contribution of induced CD signals of complex 1. Another interesting finding is that the hypochromism of the complexes is not linear relation to their DNA-binding affinities. In order to deeply study these experimental phenomena and trends, the density functional theory (DFT) and time-dependent DFT (TDDFT) computations were carried out, and on the basis of the DFT/TDDFT results and the frontier molecular orbital theory, the trend in DNA-binding affinities, the spectral properties as well as the interesting phenomena of larger extent of hypochromism but relatively smaller K(b) values for the title complexes have been reasonably explained.  相似文献   

13.
Two new complexes of Ru(III) with purine base derivatives, [mer-RuCl(3)(acv)(DMSO-S)(C(2)H(5)OH)].C(2)H(5)OH (1) (acv=acyclovir, DMSO=dimethyl sulfoxide) and [trans-RuCl(4)(guaH)(DMSO-S)].2H(2)O (2) (guaH=protonated molecule of guanine), were prepared from the same Ru(III) precursor, [trans-RuCl(4)(DMSO-S)(2)](-), by substitution of one DMSO-S. Coordination of acv induced also replacement of one chloride by an ethanol molecule. This reactivity difference was explained by striking contrasts in the hydrogen bonding schemes of the two complexes, evidenced in their X-ray crystal structures. In 1 the guanine derivative acyclovir is coordinated to ruthenium through the N(7) atom, while in 2 the protonated guanine molecule is bound through the N(9) atom. Both complexes were also characterized by various physico-chemical methods in the solid state and in the solution. In vitro, the biological activity of 2 and of the previously described complexes [mer-RuCl(3)(acv)(DMSO-S)(CH(3)OH)].0.5CH(3)OH (3) and [mer-RuCl(3)(acv)(DMSO-S)(H(2)O)].H(2)O (4) on tumour cells appear to be very similar to that of NAMI-A (NAMI-A=[ImH][trans-RuCl(4)(DMSO-S)Im]). All compounds are only weakly active on tumour cell proliferation but show an interesting proadhesive effect that suggest possible activity on tumour malignancy.  相似文献   

14.
Electrospray ionization spectra of potential cyanide-containing gold-drug metabolites revealed additional, weak, unanticipated peaks at approximately twice the mass of the gold(I) and gold(III) cyanide complexes. The exact masses correspond to proton-linked bimetallic complexes, [H[Au(CN)(m)](2)](-), (m=2,4). Further investigation revealed a total of 12 examples, including trimetallic complexes, [H(2)[Au(CN)(m)](3)](-); mixed species with two complexes, [H[Au(CN)(2)][Au(CN)(4)]](-); and thiolato species, [H[(RS)Au(CN)(3)](2)](-). trans-[AuX(2)(CN)(2)Cl(2)](-) and trans-[AuX(2)(CN)(2)Br(2)](-) generated (35)Cl/(37)Cl and (79)Br/(81)Br isotopic patterns for the protonated bi- and tri-metallic analogues which were in good agreement with the presence of four or six halide ligands, respectively. Concentration-dependent studies demonstrated that the signals are independent of the solution concentrations of mono-metallic precursors, suggesting formation in the gas phase during or following droplet desolvation.  相似文献   

15.
High performance capillary electrophoresis (HPCE) as well as high performance liquid chromatography-mass spectrometry (HPLC-MS) have been applied to the separation, identification and quantification of the tumor-inhibiting ruthenium compounds HIm trans-[RuCl4(im)2] (im = imidazole) and HInd trans-[RuCl4(ind)2] (ind = indazole) and their hydrolysis products. The half-lives for the hydrolytic decomposition of the Ru(III) compounds were determined by monitoring the relative decrease of the original complex anion under different conditions by means of capillary electrophoresis. The decomposition follows pseudo-first-order kinetics. The rate constants in water at 25 degrees C are 1.102 +/- 0.091 x 10(-5) s-1 for HIm trans-[RuCl4(im)2] and 0.395 +/- 0.014 x 10(-5) s-1 for HInd trans-[RuCl4(ind)2]. About 8% of HIm trans-[RuCl4(im)2] but only about 2% of HInd trans-[RuCl4(ind)2] were hydrolyzed after 1 h at room temperature. Whereas the hydrolysis rate of the imidazole complex is independent of the pH value, the indazole complex hydrolyzes much faster at higher pH. The half-life of HInd trans-[RuCl4(ind)2] in phosphate buffer at pH 6.0 and 37 degrees C is 5.4 h, whereas it is less than 0.5 h at pH 7.4. In contrast to the imidazole complex, where no dependence on the buffer system was observed, hydrolysis of the indazole complex is even faster if a buffer containing hydrogen carbonate is used. The formation of [RuCl2(H2O)2(im)2]+ could be demonstrated by HPLC-MS measurements. In the case of the indazole complex, a release of the indazole ligands results in the formation of [RuCl4(H2O)2]-.  相似文献   

16.
The NO donor trans-[Ru(NO)(NH(3))(4)(py)](BF(4))(3).H(2)O (py=pyridine) was loaded into poly-lactic-co-glycolic acid (PLGA) microparticles using the double emulsification technique. Scanning electron microscopy (SEM) and dynamic light scattering revealed that the particles are spherical in shape, have a diameter of 1600nm, and have low tendency to aggregate. The entrapment efficiency was 25%. SEM analysis of the melanoma cell B16-F10 in the presence of the microparticles containing the complex trans-[Ru(NO)(NH(3))(4)(py)](BF(4))(3).H(2)O (pyMP) showed that the microparticles were adhered to the cell surface after 2h of incubation. The complex with concentrations lower than 1x10(-4)M did not show toxicity in B16-F10 murine cells. The complex in solution is toxic at higher concentrations (>1x10(-3)M), with cell death attributed to NO release following the reduction of the complex. pyMP is not cytotoxic due to the lower bioavailability and availability of the entrapped complex to the medium and its reducing agents. However, pyMP is phototoxic upon light irradiation. The phototoxicity strongly suggests that cell death is due to NO release from trans-[Ru(NO)(NH(3))(4)(py)](3+). This work shows that pyMP can serve as a model for a drug delivery system carrying the NO donor trans-[Ru(NO)(NH(3))(4)(py)](BF(4))(3).H(2)O, which can release NO locally at the tumor cell by irradiation with light only.  相似文献   

17.
The interaction between hexakis(imidazole) manganese(II) terephthalate ([Mn(Im)(6)](teph).4H(2)O) and salmon sperm DNA in 0.2M pH 2.30 Britton-Robinson buffer solution was studied by fluorescence spectroscopy and cyclic voltammetry. Increasing fluorescence was observed for [Mn(Im)(6)](2+) with DNA addition, while quenching fluorescence phenomenon appeared for EB-DNA system when [Mn(Im)(6)](2+) was added. There were a couple quasi-reversible redox peaks of [Mn(Im)(6)](2+) from the cyclic voltammogram on the glassy carbon electrode. The peak current of [Mn(Im)(6)](2+) decreased with positive shift of the formal potential in the presence of DNA compared with that in the absence of DNA. All the experimental results indicate that [Mn(Im)(6)](2+) can bind to DNA mainly by intercalative binding mode. The binding ratio of the DNA-[Mn(Im)(6)](2+) association complex is calculated to be 1:1 and the binding constant is 4.44x10(3) M(-1). By using [Mn(Im)(6)](teph).4H(2)O as the electrochemical hybridization indicator, the DNA electrochemical sensor was prepared by covalent interaction and the selectivity of ssDNA modified electrode were described. The results demonstrate the use of electrochemical DNA biosensor in the determination of complementary ssDNA.  相似文献   

18.
We studied the effect of furosemide on GABAA-induced 36Cl transport and GABAA-induced Cl--ATPase activity in synaptic membranes of fish brain. At physiological pH 7.4, GABA (0.1–100 µM) stimulated 36Cl influx in synaptoneurosomes and Cl--ATPase activity in synaptic membranes. Furosemide (0.1–0.5 mM) removed the activating effect of the mediator on chloride transport and enzyme activity (I50 equaled 0.16 and 0.12 mM, respectively). In the absence of the mediator, picrotoxin (50 µM) activated the basal 36Cl influx in synaptoneurosomes and the basal Mg2+-ATPase activity of synaptic membranes. Furosemide (1 mM) removed the activating effect of picrotoxin on both biochemical processes. The obtained data demonstrated similar sensitivities of GABAA-induced transport of 36Cl in synaptoneurosomes and of GABAA-induced Cl--ATPase activity in the synaptic membranes to furosemide and indicated the involvement of the ATPase in GABAA-induced processes. The soluble ATPase, recovered by sodium deoxycholate solubilization of the membranes, remained sensitive to GABAA-ergic ligands, which suggested proximity of their binding sites with ATP hydrolysis sites in the protein molecule and their structural coupling.Translated from Izvestiya Akademii Nauk, Seriya Biologicheskaya, No. 1, 2005, pp. 18–22.Original Russian Text Copyright © 2005 by Menzikov, Menzikova.  相似文献   

19.
Attempted syntheses of ruthenium(II) monosubstituted squarate complexes in acetonitrile using cis-[RuCl2(dmso)4] and anisole-, methoxy-, methyl- and diphenylamino-squarate ligands, respectively, resulted in the formation in each case of the monomer cis, fac-Ru(CH3CN)Cl2(dmso)3 (1) with the ruthenium atom in a distorted octahedral environment. A second crop of crystals harvested from the reaction with the methoxysquarate ligand was identified as the oxalato-bridged dimer [{cis-(CH3CN)(Cl)(dmso)2Ru}2(μ-C2O4)] (2). When cis-[RuCl2(dmso)4] and methylsquarate were reacted in aqueous solution instead of acetonitrile, the dimer [{fac-(Cl)(dmso)3Ru}2(μ-C2O4)] (3) was produced. The dimers 2 and 3 are formed from oxidation/ring opening of the methoxy- and methyl-squarate ligands, respectively. Use of the salts of these ligands instead of their non-ionised forms under different reaction conditions, afforded [Na] fac-[RuCl3(dmso)3] (4) and [(C4H9)4N]2[(C4O4)(C4H2O4)2] (7), respectively, which were shown to be products of competing reactions. The information acquired from these failed attempts has provided the basis for the development of a strategy to overcome these problems and lead to a successful synthetic route to ruthenium(II) monosubstituted squarates.  相似文献   

20.
An interesting series of mononuclear organoruthenium complexes of formulation [Ru(CO)(PPh3)2(ap-R)] (where ap-R = -H, -Cl, -Me, -OMe, -OEt) have been synthesized from the reaction of five 2-(arylazo)phenol ligands with ruthenium(II) precursor [RuH(Cl)(CO)(PPh3)3] in benzene under reflux. The 2-(arylazo)phenolate ligands behave as dianionic tridentate ligand and are coordinated to ruthenium through C, N and O by dissociation of the phenolic and phenyl proton at the ortho position of the phenyl ring forming two five-membered chelate rings. These complexes have been characterized by elemental analysis, FT-IR, 1H NMR and UV-visible spectroscopy. In dichloromethane solution all the metal complexes exhibit characteristic metal-to-ligand charge transfer (MLCT) absorption and emission bands in the visible region. The structures of [Ru(CO)(PPh3)2(ap-H)] and [Ru(CO)(PPh3)2(ap-Cl)] have been determined by X-ray crystallography. Cyclic voltammetric data of all the complexes show a Ru(III)/Ru(II) oxidation and reduction Ru(II)/Ru(I) within the range 0.74-0.84 V and -0.38 to -0.50 V vs saturated calomel electrode (SCE) respectively. The potentials are observed with respect to the electronic nature of substituents (R) in the 2-(arylazo)phenolate ligands. Further, the free ligands and their ruthenium complexes have also been screened for their antibacterial and antifungal activities, which have shown great promise in inhibiting the growth of both gram +ve and gram -ve bacteria Staphylococcus aureus and Escherichia coli and fungus Candida albicans and Aspergillus niger. These results made it desirable to delineate a comparison between free ligands and their complexes.  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