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 cobalt(II)—cobalt(I) interconversion in a number of vitamin B1 2 derivatives was investigated by cyclic voltammetry. Particular attention was focused on the factors determining whether the interconversion is reversible. When the lower axial coordination position is occupied by a strong ligand, such as the benzimidazole nucleotide in “base on” cobalamins, the cobalt(II)—cobalt(I) interconversion is irreversible due to a slow reduction of the cobalt(II). However, when the lower axial coordination position is free of a strong ligand, as in most cobinamides or in “base off” cobalamins, the cobalt(II)—cobalt(I) interconversion is nearly perfectly reversible. Possible implications of the observations to B1 2-dependent enzymes are briefly discussed.  相似文献   

2.
A series of five tris(2-pyridylmethyl)amine (tpa) ruthenium complexes [Ru(tpa)(N–N)](PF6)2 with N–N = bpy (2,2′-bipyridine), phen (1,10-phenanthroline), dpq (dipyrido[3,2-d:2′,3′-f]quinoxaline), dppz (dipyrido[3,2-a;2′,3′-c]phenazine), and dppn (4,5,9,16-tetraazadibenzo[a,c]naphthacene) was prepared and characterized by NMR, UV–Visible (UV/Vis), and fluorescence spectroscopy as well as cyclic voltammetry. Structures optimized with density functional theory methods (DFT, BP86, TZVP) without constraints show C1 symmetry while in solution, the 1H and 13C NMR spectra are in accordance with an average Cs symmetry. This is thought to be due to a low energy barrier for flipping of the equatorial pyridine ring from one side of the N–N plane to the other. The electronic structure of the compounds was studied with DFT and a change in the highest occupied molecular orbital (HOMO) character from Ru t2g for the bpy, phen, and dpq to N–N ligand-based for the dppz and dppn complexes was found. TDDFT calculations showed dominant N–N-based intra-ligand charge transfer (ILCT) transitions in the latter two complexes mixed with metal-to-ligand charge transfer (MLCT) bands found for all five compounds. DNA binding of the complexes was studied with UV/Vis titrations, the fluorescent ethidium bromide displacement assay, and CD spectroscopy. The affinity increases with the aromatic surface area of of the bidentate N–N ligand in the order bpy  phen < dpq < dppz  dppn. Viscosity measurements support an intercalative binding mode for the latter three compounds, while the others did not show a pronounced effect of the hydrodynamic properties of calf thymus (CT) DNA.  相似文献   

3.
The stability constants of the ternary Cu(II), Ni(II), and Co(II) complexes containing pyridoxamine (PM) and as a second ligand (L) glycine, DL-alanine, DL-valine, and β-phenylalnine were determined by pH-metric titration in 0.50 M KNO3 at 30°C. The corresponding constants of the equilibrium, log X, are greater than would be expected for purely statistical reasons (log X = 0.6), except for few complex cases of Co(II). It has been also concluded that amino acids compete more than pyridoxamine for Ni(II) and Co(II) through the formation of 1:2:1:0 species rather than 2:1:1:0 of PM:L:M2+:H+.  相似文献   

4.
1H NMR spectroscopy was applied to study the reactions of cis-[Pd(L)(H2O)2]2+ complexes (L is en, pic and dpa) with the N-acetylated tripeptides L-methionylglycylglycine, MeCOMet–Gly–Gly, and glycyl–L-methionyl–glycine, MeCOGly–Met–Gly. All reactions were performed in the pH range 2.0–2.5 with equimolar amounts of the cis-[Pd(L)(H2O)2]2+ complex and the tripeptide at 60 °C. The hydrolytic reactions of the cis-[Pd(en)(H2O)2]2+, cis-[Pd(pic)(H2O)2]2+ and cis-[Pd(dpa)(H2O)2]2+ complexes with MeCOMet–Gly–Gly were regioselective and only the amide bond involving the carboxylic group of methionine was cleaved. However, in the reactions of these three Pd(II) complexes with MeCOGly–Met–Gly, two amide bonds, Met–Gly and MeCO–Gly, were cleaved. From UV–Vis spectrophotometry studies, it was found that the rate-determining step of these hydrolytic reactions is the monodentate coordination of the corresponding Pd(II) complex to the sulfur atom of the methionine side chain. The rate of the cleavage of these amide bonds is dependent on the nature of the bidentate coordinated diamine ligand L (en > pic > dpa). The hydrolytic reaction of cis-[Pd(L)(H2O)2]2+-type complexes with MeCOMet–Gly–Gly, containing the methionine side chain in the terminal position of the peptide, is regioselective while in the reaction of these Pd(II) complexes with MeCOGly–Met–Gly, none selective cleavage of the peptide occurs. This study contributes to a better understanding of the selective cleavage of methionine-containing peptides employing palladium(II) complexes as catalysts.  相似文献   

5.
Analogues of cytotoxic cis and trans dichloridoplatinum(II) complexes with one ammonia and one aromatic amine (cis- and trans-[PtCl2(aromatic amine)(NH3)]) were synthesised in which the aromatic group was replaced by the fluorescent ligand 7-azaindole (1). Coordination resulted in almost complete quenching of the fluorescence and the ligand had a effect on the biological activities of the cis and trans isomers similar to that previously reported for aromatic amines as is exemplified by them having similar cytotoxicities (IC50 3.6(5) and 6.0(19) μM, respectively). Observation of fluorescence following treatment of the cis complex with cysteine, glutathione, or methionine suggests labilisation and subsequent loss of the putative non-leaving group ligands. No such effect was observed for the trans complex which does not experience trans labilisation. Two-photon excitation of cells that had been treated with the complexes gave rise to observable fluorescence, suggesting ligand displacement for both complexes. The fluorescence appears to be localised in the lysosomes or late endosomes. These complexes are excellent models of analogues of cytotoxic cis and trans complexes with aromatic amine ligands and can be used to study the metabolism of the non-leaving group positions.  相似文献   

6.
One subclass of B12-requiring enzymes is now known to bind their B12 coenzymes “base-off,” with a histidine residue from the protein supplying an imidazole ligand to the cobalt center. Recent results from Sirovatka and Finke (J.M. Sirovatka and R.G. Finke, J.Am. Chem. Soc. 119, (1997) 3057) show that imidazole has an extraordinary trans effect on the mode of carbon–cobalt bond cleavage in coenzyme B12 analogs, compared to pyridine or the natural 5,6-dimethylbenzimidazole ligand, and it was suggested that a differential steric effect could, in part, account for the uniqueness of the imidazole ligand. Such a differential steric effect for imidazole and pyridine is now demonstrated by studies of the thermodynamics of ligation of these ligands to the α and β diastereomers of two alkylcobinamides (RCbi+s, derivatives of cobalamins which lack the normal axial nucleotide) based on the known differences in steric crowding of the α (“lower”) and β (“upper”) axial ligand positions of cobalt corrinoids. Imidazole binds more tightly than pyridine to both diastereomers of NCCH2Cbi+ and CF3Cbi+, in all cases due to a more favorable entropy change, which is the result of lowered steric interference with corrin side chain thermal motions.  相似文献   

7.
Recently we have found that the metallocarbonyl complexes (η5-C5H5)M(CO)x(η1-N-maleimidato) (M = Fe, Mo, W; x = 2 or 3) bearing a maleimide function were irreversible inhibitors of the enzyme papain. To get further insight into the binding mechanism of these compounds we synthesized the related complexes (η5-C5H5)M(CO)x(η1-N-succinimidato) (M = Fe, Mo, W; x = 2 or 3) that lacked the ethylenic bond responsible for alkylation of the cysteine 25 thiol group in the papain‘s catalytic pocket. We performed kinetic studies of the interaction of the synthesized complexes towards papain. We found that they act as reversible inhibitors of the enzyme with IC50 values in the range 480–1700 μM. Docking experiments confirmed binding of these complexes to the enzyme’s catalytic pocket.  相似文献   

8.
Antimutagenic activity of aqueous extracts of the South African herbal teas, Aspalathus linearis (rooibos) and Cyclopia spp. (honeybush) was compared with that of Camellia sinensis (black, oolong and green) teas in the Salmonella mutagenicity assay using aflatoxin B1 (AFB1) and 2-acetylaminofluorene (2-AAF) as mutagens. The present study presents the first investigation on antimutagenic properties of C. subternata, C. genistoides and C. sessiliflora. The herbal teas demonstrated protection against both mutagens in the presence of metabolic activation, with the exception of “unfermented” (green/unoxidised) C. genistoides against 2-AAF, which either protected or enhanced mutagenesis depending on the concentration. Antimutagenic activity of “fermented” (oxidised) rooibos was significantly (P < 0.05) less than that of Camellia sinensis teas against AFB1, while for 2-AAF it was less (P < 0.05) than that of black tea and similar (P > 0.05) to that of oolong and green teas. Antimutagenic activity of unfermented C. intermedia and C. subternata exhibited a similar protection as fermented rooibos against AFB1. Against 2-AAF, fermented rooibos exhibited similar protective properties than unfermented C. intermedia and C. sessiliflora. Unfermented rooibos was less effective than the C. sinensis teas and fermented rooibos, but had similar (P > 0.05) antimutagenicity to that of fermented C. sessiliflora against AFB1 and fermented C. subternata against 2-AAF. Fermented C. intermedia and C. genistoides exhibited the lowest protective effect against 2-AAF, while fermented C. intermedia exhibited the lowest protection when utilising AFB1 as mutagen. Aspalathin and mangiferin, major polyphenols in rooibos and Cyclopia spp., respectively, exhibited weak to moderate protective effects when compared to the major green tea catechin, (−)epigallocatechin gallate (EGCG). Antimutagenic activity of selected herbal tea phenolic compounds indicated that they contribute towards (i) observed antimutagenic activity of the aqueous extracts against both mutagens and (ii) enhancement of the mutagenicity of 2-AAF by unfermented C. genistoides. Antimutagenic activity of the South African herbal teas was mutagen-specific, affected by fermentation and plant material, presumably due to changes and variation in phenolic composition.  相似文献   

9.
In order to systematically perform an experimental and theoretical study on DNA binding and photocleavage properties of transition metal complexes of the type [M(L)2(L1)](PF6)n · xH2O (where M = Co(III) or Ni(II), L = 1,10-phenanthroline or 2.2′ bipryidine, L1 = Thiophene [2,3-b] quinoline (qt), n = 3 or 2 and x = 5 or 2) have been synthesized and characterized by elemental analysis, IR, 1H NMR, UV and magnetic susceptibility data. The DNA-binding properties of these complexes have been investigated with UV-Vis, viscosity measurements, thermal denaturation and cyclic voltametric studies. It is experimentally found that all the complexes are bound to DNA via intercalation in the order [Co(bpy)2(qt)](PF6)3 > [Co(phen)2(qt)](PF6)3 > [Ni(phen)2(qt)](PF6)2 > [Ni(bpy)2(qt)](PF6)2. The photocleavage studies with pUC19 DNA shows that all these complexes promoted the conversion of SC form to NC form in absence of ‘inhibitors’.  相似文献   

10.
Intramolecular M(II)H–C interactions (M(II)=Cu(II), Pd(II)) involving a side chain alkyl group of planar d8 and d9 metal complexes of the N-alkyl (R) derivatives of N,N-bis(2-pyridylmethyl)amine with an N3Cl donor set were established by structural and spectroscopic methods. The methyl group from the branched alkyl group (R = 2,2-dimethylpropyl and 2-methylbutyl) axially interacts with the metal ion with the MC and MH distances of 3.056(3)–3.352(9) and 2.317(1)–2.606(1) Å, respectively, and the M–H–C angles of 122.4–162.3°. The Cu(II) complexes showing the interaction have a higher redox potential as compared with those without it, and the 1H NMR signals of the interacting methyl group in Pd(II) complexes shifted downfield relative to the ligand signals. Dependence of the downshift values on the dielectric constants of the solvents used indicated that the M(II)H–C interaction is mainly electrostatic in nature and may be regarded as a weak hydrogen bond. Implications for possible environmental effects of the leucine alkyl group at the type 1 Cu site of fungal laccase are also discussed.  相似文献   

11.
Kinetics for the breakdown of the trinuclear chromium acetate cluster with a series of monoprotic and diprotic amino acid ligands and with glutathione in aqueous media have been investigated spectrophotometrically at pH 3.5–5.5 and in a temperature range of 45–60 °C. Under pseudo-first-order conditions, reactions with these ligands exhibited biphasic kinetic behavior that can be accounted for by a consecutive two-step reaction, A → B → C, where A is assumed to be a forced ion pair, B an intermediate and C is the product; experimental data fit to a biexponential equation for the transformation. Rates for kshort, klong, and kobs were determined by manual extrapolation of absorbance data or curve-fitting routines; associated activation parameters for each step of the reaction were calculated using the Eyring equation. Rates for the first and second steps of the reaction are on the order of 10−4 and 10−5 s−1, respectively. The large negative values of ΔS and smaller ΔH in the first step indicate an associative step, while high positive values of ΔS in the second step indicate dissociation. To account for the results mechanistically, the results are interpreted to be a first step of ligand exchange with a pseudo-axial aqua ligand, followed by a dissociative step involving acetate or oxo ligand displacement. The dissociative step is the rate determining step, with kobs ≈ klong.The results demonstrate reaction pathways that are available to the Cr(III) metal centers that may be physiologically relevant in the ligand-rich environment of biological systems. Under general conditions Cr(III) clusters may be expected to be broken down, unless some unique biological environment stabilizes the cluster. The present study has application to the processes related to Cr(III) transport and excretion, to potential mechanisms of Cr(III) action in a biological setting, and to the pharmacokinetics of Cr(III) supplements for animal and human consumption.  相似文献   

12.
A modified β-cyclodextrin bearing a 2-aminomethylpyridine binding site for copper(II) (6-deoxy-6-[N-(2-methylamino)pyridine)]-β-cyclodextrin, CDampy was synthesized by C6-monofunctionalization. The acid-base properties of the new ligand in aqueous solution were investigated by potentiometry and calorimetry, and its conformations as a function of pH were studied by NMR and circular dichroism (c.d.). The formation of binary copper(II) complexes was studied by potentiometry, EPR, and c.d. The copper(II) complex was used as chiral selector for the HPLC enantiomeric separation of underivatized aromatic amino acids. Enantioselectivity in the overall stability constants of the ternary complexes with D- or L-Trp was detected by potentiometry, whereas the complexes of the Ala enantiomers did not show any difference in stability. These results were consistent with a preferred cis coordination of the amino group of the ligand and of the amino acid in the ternary complexes (“cis effect”), which leads to the inclusion of the aromatic side chain of D-Trp, but not of that of L-Trp. In Trp-containing ternary complexes, the two enantiomers showed differences in the fluorescence lifetime distribution, consistent with only one conformer of D-Trp and two conformers of L-Trp, and the latter were found to be more accessible to fluorescence quenching by acrylamide and KI. Chirality 9:341–349, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

13.
In order to study the influence of nutrients on the growth characteristics of the dominant dinoflagellates, Ceratium furca and Ceratium fusus, in the temperate coastal area of Sagami Bay, Japan, we conducted field monitoring from January 2000 to December 2005 and performed laboratory culture experiments. In the field study, population densities of C. furca and C. fusus were high, even in low nutrient concentrations (N: 1.58 μM, P: 0.17 μM). Both species were more abundant in the surface and sub-surface layers than in the bottom layers during the stratification periods. In the laboratory study, the specific growth rates of C. furca and C. fusus increased gradually along with increasing nutrients up to the T5 (N: 5 μM, P: 0.5 μM) and T10 (N: 10 μM, P: 1 μM) concentration levels, after which the growth rate plateaued at the T50 (N: 50 μM, P: 5 μM) concentration level. In contrast, the nutrient uptake rates of both species continuously increased, indicating “luxury consumption”, i.e., excessive cellular storage not related to growth rate. The half-saturation constants (Ks) of C. furca for nitrate (0.49 μM) and phosphate (0.05 μM) were slightly higher than C. fusus (0.32 and 0.03 μM, respectively). We offer two reasons why the two Ceratium population densities were maintained at high levels in low nutrient conditions. First, these two species have a competitive advantage over other algal species because of low Ks values and specific characteristics for nutrient uptake such as luxury consumption. Their ability to obtain nutrients through alternative methods, such as phagotrophy, might contribute to bloom formation and population persistence. Second, the cell densities of both Ceratium species increased along with nitrate concentrations in the media even when phosphorus was held constant. In particular, the growth of C. furca was directly supported by various nitrogen sources such as nitrate, ammonium, and urea, although the highest growth rates were observed only in the nitrate-enriched cultures. Our field and laboratory results revealed that the growth rates of the two Ceratium species increased readily in high N:P nutrient conditions (i.e., conditions of P limitation) indicating an advantage over other algal species in phosphorus-limited environments such as Sagami Bay.  相似文献   

14.
The reactions of Pd(II) and Pt(II) with 2-Acetyl Pyridine N(4)-Ethyl-Thiosemicarbazones, HAc4Et and 2-Acetyl Pyridine N(4)-1-(2-pyridyl)-piperazinyl Thiosemicarbazone, HAc4PiPiz and 2-Formyl Pyridine N(4)-1-(2-pyridyl)-piperazinyl Thiosemicarbazone, HFo4PiPiz afforded the complexes, [Pd(Ac4Et)], 1, [Pd(HAc4Et)2]Cl2, 2 and [Pd(Ac4Et)2], 3[Pt(Ac4Et)], 4, [Pt(HAc4Et)2]Cl2, 5, [Pt(Ac4Et)2], 6 and [Pd(Fo4PipePiz)Cl], 7, [Pd(Fo4PipePiz)2], 8, [Pd(Ac4PipePiz)Cl], 9 and [Pd(Ac4PipePiz)2], 10. The crystal structure of the complex [Pt(Ac4Et)2], 6 has been solved. The platinum(II) atom is in a square planar environment surrounded by two cis nitrogen atoms and two cis sulfur atoms. The ligands are not equivalent, one being tridentate with (N,N,S) donation, the other being monodentate using only the sulfur atom to coordinate to the metal. The tridentate ligand shows a Z, E, Z configuration while the monodentate ligand shows an E, E, Z. Inter-molecular hydrogen bonds stabilize the structure, while the crystal packing is determined by –, and Pt – C interactions. The antibacterial effect of Pd(II) and Pt(II) complexes were studied in vitro. The complexes were found to have effect on Gram(+) bacteria, while the same complexes showed no bactericidal effect on Gram(–) bacteria. The effect of the Pd(II) and Pt(II) complexes on the in vitro DNA strand breakage was studied by agarose gel electrophoresis. The complexes 1-6 were found to exhibit a cytotoxic potency in a very low micromolar range and to be able to overcome the cisplatin resistance of A2780/Cp8 cells (Kovala-Demertzi et al. 2000).  相似文献   

15.
The condensation of (1H-benzimidazole-2-yl) methanamine, with 2-hydroxy naphthaldehyde lead to Schiff base ligand (H2L) (1). This was later reacted with metal salts (ZnCl2, CrCl3·6H2O, and MnCl2·4H2O) to afford the corresponding metal complexes. Biological activity findings indicate that the metal complexes have promising activity against Escherichia coli and Bacillus subtilis and modest activity against Aspergillus niger. The in vitro anticancer activities of Zn (II), Cr (III), and Mn (II) complexes were investigated and the best results were observed with Mn (II) complex as the most potent cytotoxic agent toward human cell lines colorectal adenocarcinoma HCT 116, hepatocellular carcinoma HepG2 and breast adenocarcinoma MCF-7 with 0.7, 1.1 and 6.7 μg of inhibitory concentration IC50 values respectively. Consequently, the Mn (II) complex and ligand were docked inside the energetic site of ERK2 and exhibited favorable energy for binding. The investigation of biological tests towards mosquito larvae indicates that Cr (III) and Mn (II) complexes manifest strong toxicity against Aedes aegypti larvae with 3.458 and 4.764 ppm values of lethal concentration LC50, respectively.  相似文献   

16.
Sulfated polysaccharides were localized in the cuticle, cortex and medulla of the gametophyte thallus, being more concentrated in the intercellular matrix than in the cell walls. During the water extraction sequence, a small percentage of galactan sulfates (5.1% of dry seaweed) with average low Mr (6–11.4 kDa) were extracted at room temperature without disturbing the cellular arrangement, while sulfated galactans of average medium Mr (18–45 kDa) were obtained by further hot-water extractions (52.4% of dry seaweed), with diorganization of the tissue. The residue (40.0% of dry seaweed) still contained carrageenan-type (major) and agaran-type (minor) galactans. Part of these galactans was extracted with 8.4% LiCl solution in DMSO, from which “pure” κ/ι-carrageenans were isolated.Carrageenans and agarans were extracted in a ratio 1:0.5, showing the highest amount of agaran-structures for a carrageenophyte. The galactans comprise alternating 4-sulfated (major) and non-sulfated (minor) 3-linked β-d-galactopyranose units, and 4-linked α-galactopyranose units with the following substitutions: (i) non-sulfated and 2-sulfated 3,6-anhydro-α-d-galactopyranose residues in the carrageenan-structures, which belong to the κ-family (κ/ι-carrageenans); (ii) 3-sulfated α-l-galactopyranose units and 2-sulfated 3,6-anhydro-α-l-galactopyranose residues in the agaran-structures.Alkaline treatment and alkaline dialysis of the main extracts gave “pure” κ/ι-carrageenans, showing that carrageenan molecules are extracted together with low Mr agarans or agaran-dl-hybrids.  相似文献   

17.
Antifungal effectivity and utility of cinnamaldehyde is limited because of its high MIC and skin sensitivity. In this study, α-methyl trans cinnamaldehyde, a less irritating derivative, have been self coupled and complexed with Co(II) and Ni(II) to generate N, N′–Bis (α-methyl trans cinnamadehyde) ethylenediimine [C22H24N2], [Co(C44H48N4)Cl2] and [Ni(C44H48N4)Cl2]. Ligand and complexes were characterized on the basis of FTIR, ESI–MS, IR and 1HNMR techniques. Synthesized ligand [L] and complexes were investigated for their MICs, inhibition of ergosterol biosynthesis and H+ extrusion against three strains of Candida: C. albicans 44829, C. tropicalis 750 and C. krusei 6258. Average of three species MIC of methyl cinnamaldehyde is 317 μg/ml (2168 μM). Compared to methyl cinnamaldehyde ligand [L], Co(II) and Ni(II) complex are found to be 4.48, 17.78 and 21.46 times more effective in liquid medium and 2.73, 8.93 and 10.38 times more effective in solid medium. At their respective MIC90 average inhibition of ergosterol biosynthesis caused by methyl cinnamaldehyde, ligand [L], Co(II) and Ni(II) complex, respectively was 80, 78, 90 and 93%. H+ extrusion was also significantly inhibited but did not co-relate well with MIC90. Results indicate ergosterol biosynthesis as site of action of α-methyl cinnamaldehyde, synthesized ligand and complexes. α-methyl cinnamaldehyde and ligand did not show any toxicity against H9c2 rat cardiac myoblast cell, whereas Co(II) and Ni(II) complexes on an average produced 19% cellular toxicity.  相似文献   

18.
Cobalt(II), nickel(II), copper(II) and zinc(II) complexes with 2-acetylthiophene benzoylhydrazone have been synthesized and characterized by elemental analyses, magnetic susceptibility measurements, electronic, IR, NMR and ESR spectral techniques. The molecular structures of ligand and its copper(II) complex have been determined by single crystal X-ray diffraction technique. The Cu(II) complex possesses a CuN2O2 chromophore with a considerable delocalization of charge. The structure of the complex is stabilized by intermolecular π–π stacking and C–H?π interactions. Hatbh acts as a monobasic bidentate ligand in all the complexes bonding through a deprotonated C–O and >CN groups. Electronic spectral studies indicate an octahedral geometry for the Ni(II) complex while square planar geometry for the Co(II) and Cu(II) complexes. ESR spectrum of the Cu(II) complex exhibits a square planar geometry in solid and in DMSO solution. The trend g|| > g > 2.0023 indicates the presence of an unpaired electron in the dx2-y2 orbital of Cu(II). The electro-chemical study of Cu(II) complex reveals a metal based reversible redox behavior. The Ni(II) complex shows exothermic multi-step decomposition pattern of the bonded ligand. The ligand and its most of the metal complexes show appreciable corrosion inhibition properties for mild steel in 1 M HCl medium. [Co(atbh)2] complex exhibited the greatest impact on corrosion inhibition among the other compounds.  相似文献   

19.
Different chemical extractants (NaCl, EDTA, HCl and NaOH) and physical methods (ultrasonication and heating) were examined by their efficacies of extracting “attached” exopolymeric substances (EPS) secreted by marine bacterium Sagittula stellata (SS) and terrestrial bacterium Pseudomonas fluorescens Biovar II (PF). Extraction by 0.5 N HCl for 3 h was best for SS while extraction by 0.05 N NaCl for 3–5 h was regarded as optimal for PF. Improvements in EPS purification included a pre-diafiltration step to remove the broth material and reduce the solution volume, thus the usage of ethanol, and time. The EPS harvested at the optimal time and purified by the improved method were enriched in polysaccharides, with smaller amounts of proteins, thus having amphiphilic properties. Isoelectric focusing of 234Th or 240Pu labeled EPS showed both actinides were strongly bound to macromolecules with low pI, similar to reported marine or soil colloidal natural organic matter (NOM).  相似文献   

20.
Protein binding, DNA binding/cleavage and in vitro cytotoxicity studies of 2-((3-(dimethylamino)propyl)amino)naphthalene-1,4-dione (L) and its four coordinated M(II) complexes [M(II) = Co(II), Cu(II), Ni(II) and Zn(II)] have been investigated using various spectral techniques. The structure of the ligand was confirmed by spectral and single crystal XRD studies. The geometry of the complexes has been established using analytical and spectral investigations. These complexes show good binding tendency to bovine serum albumin (BSA) exhibiting high binding constant values (105 M?1) when compared to free ligand. Fluorescence titration studies reveal that these compounds bind strongly with CT-DNA through intercalative mode (Kapp 105 M?1) and follow the order: Cu(II) > Zn(II) > Ni(II) > Co(II) > L. Molecular docking study substantiate the strength and mode of binding of these compounds with DNA. All the complexes efficiently cleaved pUC18-DNA via hydroxyl radical mechanism and the Cu(II) complex degraded the DNA completely by converting supercoiled form to linear form. The complexes demonstrate a comparable in vitro cytotoxic activity against two human cancer cell lines (MCF-7 and A-549), which is comparable with that of cisplatin. AO/EB and DAPI staining studies suggest apoptotic mode of cell death, in these cancer cells, with the compounds under investigation.  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