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 171 毫秒
1.
Ghosh D  Lee KH  Demeler B  Pecoraro VL 《Biochemistry》2005,44(31):10732-10740
Investigators have studied how proteins enforce nonstandard geometries on metal centers to assess the question of how protein structures can define the coordination geometry and binding affinity of an active-site metal cofactor. We have shown that cysteine-substituted versions of the TRI peptide series [AcG-(LKALEEK)(4)G-NH(2)] bind Hg(II) and Cd(II) in geometries that are different from what is normally found with thiol ligands in aqueous solution. A fundamental question has been whether this structural perturbation is due to protein influence or a change in the metal geometry preference. To address this question, we have completed linear free-energy analyses that correlate the association of three-stranded coiled coils in the absence of a metal with the binding affinity of the peptides to the heavy metals, Hg(II) and Cd(II). In this paper, six new members of this family have been synthesized, replacing core leucine residues with smaller and less hydrophobic residues, consequently leading to varying degrees of self-association affinities. At the same time, studies with some smaller and longer sequenced peptides have also been examined. All of these peptides are seen to sequester Hg(II) and Cd(II) in an uncommon trigonal environment. For both metals, the binding is strong with micromolar dissociation constants. For binding of Hg(II) to the peptides, the dissociation constants range from 2.4 x 10(-)(5) M for Baby L12C to 2.5 x 10(-)(9) M for Grand L9C for binding of the third thiolate to a linear Hg(II)(pep)(2) species. The binding of Hg(II) to the peptide Grand L9C is similar in energetics for metal binding in the metalloregulatory protein, mercury responsive (merR), displaying approximately 50% trigonal Hg(II) formation at nanomolar metal concentrations. Approximately, 11 kcal/mol of the Hg(II)(Grand L9C)(3)(-) stability is due to peptide interactions, whereas only 1-4 kcal/mol stabilization results from Hg(II)(RS)(2) binding the third thiolate ligand. This further validates the hypothesis that the favorable tertiary interactions in protein systems such as merR go a long way in stabilizing nonnatural coordination environments in biological systems. Similarly, for the binding of Cd(II) to the TRI family, the dissociation constants range from 1.3 x 10(-)(6) M for Baby L9C to 8.3 x 10(-)(9) M for TRI L9C, showing a similar nature of stable aggregate formation.  相似文献   

2.
Complex formation between heavy metal ions and glutathione (GSH) is considered as the initial step in many detoxification processes in living organisms. In this study the structure and coordination between the cadmium(II) ion and GSH were investigated in aqueous solutions (pH 7.5 and 11.0) and in the solid state, using a combination of spectroscopic techniques. The similarity of the Cd K-edge and L3-edge X-ray absorption spectra of the solid compound [Cd(GS)(GSH)]ClO4·3H2O, precipitating at pH 3.0, with the previously studied cysteine compound {Cd(HCys)2·H2O}2·H3O+·ClO4 ? corresponds to Cd(S–GS)3O (dominating) and Cd(S–GS)4 four-coordination within oligomeric complexes with mean bond distances of 2.51 ± 0.02 Å for Cd–S and 2.24 ± 0.04 Å for Cd–O. For cadmium(II) solutions (C Cd(II) ~ 0.05 M) at pH 7.5 with moderate excess of GSH (C GSH/C Cd(II) = 3.0–5.0), a mix of Cd(S–GS)3O (dominating) and Cd(S–GS)4 species is consistent with the broad 113Cd NMR resonances in the range 632–658 ppm. In alkaline solutions (pH 11.0 and C GSH/C Cd(II) = 2.0 or 3.0), two distinct peaks at 322 and 674 ppm are obtained. The first peak indicates six-coordinated mononuclear and dinuclear complexes with CdS2N2(N/O)2 and CdSN3O2 coordination in fast exchange, whereas the second corresponds to Cd(S–GS)4 sites. At high ligand excess the tetrathiolate complex, Cd(S–GS)4, characterized by a sharp δ(113Cd) NMR signal at 677 ppm, predominates. The average Cd–S distance, obtained from the X-ray absorption spectra, varied within a narrow range, 2.49–2.53 Å, for all solutions (pH 7.5 and 11.0) regardless of the coordination geometry.  相似文献   

3.
4.
The present study relates to the use of cyanobacterium Nostoc muscorum as a model system for removal of heavy metals such as Pb and Cd from aquatic systems. The effects of various physicochemical factors on the surface binding and intracellular uptake of Pb and Cd were studied to optimize the metal removal efficiency of the living cells of N. muscorum. Results demonstrated that a significant proportion of Pb and Cd removal was mediated by surface binding of metals (85 % Pb and 79 % Cd), rather than by intracellular accumulation (5 % Pb and 4 % Cd) at the optimum level of cyanobacterial biomass (2.8 g L?1), metal concentration (80 μg mL?1), pH (pH 5.0–6.0), time (15–30 min), and temperature (30–40 °C). N. muscorum has maximum amounts of metal removal (q max) capacity of 833 and 666.7 mg g?1 protein for Pb and Cd, respectively. The kinetic parameters of metal binding revealed that adsorption of Pb and Cd by N. muscorum followed pseudo-second-order kinetics, and the adsorption behavior was better explained by both Langmuir and Freundlich isotherm models. The surface binding of both the metals was apparently facilitated by the carboxylic, hydroxyl, and amino groups as evident from Fourier transform infrared spectra.  相似文献   

5.
Hepcidin is a peptide hormone that regulates the homeostasis of iron metabolism. The N-terminal domain of hepcidin is conserved amongst a range of species and is capable of binding CuII and NiII through the amino terminal copper–nickel binding motif (ATCUN). It has been suggested that the binding of copper to hepcidin may have biological relevance. In this study we have investigated the binding of CuII with model peptides containing the ATCUN motif, fluorescently labelled hepcidin and hepcidin using MALDI-TOF mass spectrometry. As with albumin, it was found that tetrapeptide models of hepcidin possessed a higher affinity for CuII than that of native hepcidin. The log K 1 value of hepcidin for CuII was determined as 7.7. CuII binds to albumin more tightly than hepcidin (log K 1 = 12) and in view of the serum concentration difference of albumin and hepcidin, the bulk of kinetically labile CuII present in blood will be bound to albumin. It is estimated that the concentration of CuII-hepcidin will be less than one femtomolar in normal serum and thus the binding of copper to hepcidin is unlikely to play a role in iron homeostasis. As with albumin, small tri and tetra peptides are poor models for the metal binding properties of hepcidin.  相似文献   

6.
The structure and speciation of the complexes formed between mercury(II) ions and glutathione (GSH = L-glutamyl-L-cysteinyl-glycine) have been studied for a series of alkaline aqueous solutions (\( C_{{{\text{Hg}}^{{2 + }}}}\,{\sim18\,{\rm{mmol}}\,{\rm{{dm^{-3}}}}}\) and C GSH = 40–200 mmol dm?3 at pH ~10.5) by means of extended X-ray absorption fine structure (EXAFS) and 199Hg NMR spectroscopy at ambient temperature. The dominant complexes are [Hg(GS)2]4? and [Hg(GS)3]7?, with mean Hg–S bond distances of 2.32(1) and 2.42(2) Å observed in digonal and trigonal Hg–S coordination, respectively. The proportions of the Hg2+–glutathione complexes were evaluated by fitting linear combinations of model EXAFS oscillations representing each species to the experimental EXAFS spectra. The [Hg(GS)4]10? complex, with four sulfur atoms coordinated at a mean Hg–S bond distance of 2.52(2) Å, is present in minor amounts (<30%) in solutions containing a large excess of glutathione (C GSH ≥ 160 mmol dm?3). Comparable alkaline mercury(II) cysteine (H2Cys) solutions were also investigated and a reduced tendency to form higher complexes was observed, because the deprotonated amino group of Cys2? allows the stable [Hg(S,N-Cys)2]2? chelate to form. The effect of temperature on the distribution of the Hg2+–glutathione complexes was studied by comparing the EXAFS spectra at ambient temperature and at 25 K of a series of glycerol/water (33/67, v/v) frozen glasses with \( C_{{{\text{Hg}}^{{2 + }} }} \,{\sim7\,{\rm{mmol}}\,{\rm{{dm^{-3}}}}} \) and C GSH = 16–81 mmol dm?3. Complexes with high Hg–S coordination numbers, [Hg(GS)3]7? and [Hg(GS)4]10?, became strongly favored when just a moderate excess of glutathione (C GSH ≥28 mmol dm?3) was used in the glassy samples, as expected for a stepwise exothermic bond formation. Addition of glycerol had no effect on the Hg(II)–glutathione speciation, as shown by the similarity of the EXAFS spectra obtained at room temperature for two parallel series of Hg(II)-glutathione solutions with \( C_{{{\text{Hg}}^{{2 + }} }} \,{\sim7\,{\rm{mmol}}\,{\rm{{dm^{-3}}}}},\) with and without 33% glycerol. Also, the 199Hg NMR chemical shifts of a series of ~18 mmol dm?3 mercury(II) glutathione solutions with 33% glycerol were not significantly different from those of the corresponding series in aqueous solution.  相似文献   

7.
HeLa S3 cells were exposed to varied concentrations of methylmercury over varied periods of time and its binding by the cells was studied using 203Hg-labeled methylmercuric chloride as radioactive marker. Also studied was the effect of cell-bound methylmercury on DNA replication and protein synthesis and on the growth rate of the cells. The results show that methylmercury binding is a rapid process, with much of the organomercurial bound within the the first 60 min of incubation, and that considerable quantities of organic mercury become affixed to the cells. The amounts of bound methylmercury, [CH3Hg(II)]bound, given in mol/cell, range from 2 × 10?16 (at 1 h of incubation and at 1 μM CH3Hg(II) in the medium) to almost 4 × 10?14 (at 24 h of incubation and at 100 μM CH3Hg(II) in the medium). A [CH3Hg(II)]bound value of about 30 × 10?16 mol/cell appears to be the threshold below which cells display a normal growth pattern and below which metabolic events such as DNA replication or protein synthesis are affected only to a minor degree but above which major changes in cell metabolism and cell growth take place. Methylmercury binding by the cells is tight so that only 20% of the bound material is released from the cells over a 3-h incubation period when the cells are placed into fresh, methylmercury-free growth medium. Analysis of the binding data in terms of binding to identical and completely independent sites yields an association constant K of 7.92 × 104 l/mol and for the maximum concentration of cellular binding sites the value 2.40 × 10?14 mol/cell or 1.45 × 1010 sites/cell. Evidence is presented which shows that cellular sulfhydryl groups do not suffice to provide all the sites taken up by methylmercury and that binding, in all likelihood, involves basic nitrogen, too. The levels of cell-bound methylmercury are such that binding to HeLa DNA and HeLa chromatin, for instance, can readily take place. Methylmercury binding data obtained by using the technique of particle-induced X-ray emission (PIXE) are in good agreement with the data obtained via isotope dilution.  相似文献   

8.
Mercuric contamination of aqueous cultures results in impairment of viability of photosynthetic bacteria primarily by inhibition of the photochemistry of the reaction center (RC) protein. Isolated reaction centers (RCs) from Rhodobacter sphaeroides were exposed to Hg2+ ions up to saturation concentration (~?103 [Hg2+]/[RC]) and the gradual time- and concentration-dependent loss of the photochemical activity was monitored. The vast majority of Hg2+ ions (about 500 [Hg2+]/[RC]) had low affinity for the RC [binding constant Kb?~?5 mM?1] and only a few (~?1 [Hg2+]/[RC]) exhibited strong binding (Kb?~?50 μM?1). Neither type of binding site had specific and harmful effects on the photochemistry of the RC. The primary charge separation was preserved even at saturation mercury(II) concentration, but essential further steps of stabilization and utilization were blocked already in the 5 < [Hg2+]/[RC]?<?50 range whose locations were revealed. (1) The proton gate at the cytoplasmic site had the highest affinity for Hg2+ binding (Kb?~?0.2 μM?1) and blocked the proton uptake. (2) Reduced affinity (Kb?~?0.05 μM?1) was measured for the mercury(II)-binding site close to the secondary quinone that resulted in inhibition of the interquinone electron transfer. (3) A similar affinity was observed close to the bacteriochlorophyll dimer causing slight energetic changes as evidenced by a?~?30 nm blue shift of the red absorption band, a 47 meV increase in the redox midpoint potential, and a?~?20 meV drop in free energy gap of the primary charge pair. The primary quinone was not perturbed upon mercury(II) treatment. Although the Hg2+ ions attack the RC in large number, the exertion of the harmful effect on photochemistry is not through mass action but rather a couple of well-defined targets. Bound to these sites, the Hg2+ ions can destroy H-bond structures, inhibit protein dynamics, block conformational gating mechanisms, and modify electrostatic profiles essential for electron and proton transfer.  相似文献   

9.
Removal of Pb(II) from an aqueous environment using biosorbents is a cost-effective and environmentally benign method. The biosorption process, however, is little understood for biosorbents prepared from plant materials. In this study, the biosorption process was investigated by evaluating four adsorption models. A fixed-bed column was prepared using a biosorbent prepared from the aquatic plant Hydrilla verticillata. The effect of bed height and flow rate on the biosorption process was investigated. The objective of the study was to determine the ability of H. verticillata to biosorb Pb(II) from an aqueous environment and to understand the process, through modeling, to provide a basis to develop a practical biosorbent column. Experimental breakthrough curves for biosorption of 50 mg L?1 aqueous Pb(II) using a fixed-bed column with 1.00 cm inner diameter were fitted to the Thomas, Adams-Bohart, Belter, and bed depth service time (BDST) models to investigate the behavior of each model according to the adsorption system and thus understand the adsorption mechanism. Model parameters were evaluated using linear and nonlinear regression methods. The biosorbent removed 65% (82.39 mg g?1 of biosorbent) of Pb(II) from an aqueous solution of Pb(NO3)2 at a flow rate of 5.0 ml min?1 in a 10 cm column. Na2CO3 was used to recover the adsorbed Pb(II) ions as PbCO3 from the biosorbent. The Pb(II) was completely desorbed at a bed height of 10.0 cm and a flow rate of 5.0 ml min?1. Fourier transform infrared (FT-IR) analysis of the native biosorbent and Pb(II)-loaded biosorbent indicated that the hydroxyl groups and carboxylic acid groups were involved in the metal bonding process. The FT-IR spectrum of Pb(II)-desorbed biosorbent showed an intermediate peak shift, indicating that Pb(II) ions were replaced by Na+ ions through an ion-exchange process. Of the four models tested, the Thomas and BDST models showed good agreement with experimental data. The calculated bed sorption capacity N0 and rate constant ka were 31.7 g L?1 and 13.6 × 10?4 L mg?1 min?1 for the Ct/C0 value of 0.02. The BDST model can be used to estimate the column parameters to design a large-scale column.  相似文献   

10.
In the early nineties, Undaria pinnatifida has been accidentally introduced to Nuevo Gulf (Patagonia, Argentina) where the environmental conditions would have favored its expansion. The effect of the secondary treated sewage discharge from Puerto Madryn city into Nueva Bay (located in the western extreme of Nuevo Gulf) is one of the probable factors to be taken into account. Laboratory cultures of this macroalgae were conducted in seawater enriched with the effluent. The nutrients (ammonium, nitrate and phosphate) uptake kinetics was studied at constant temperature and radiation (16?°C and 50 μE m?2 s?1 respectively). Uptake kinetics of both inorganic forms of nitrogen were described by the Michaelis–Menten model during the surge phase (ammonium: V max sur: 218.1 μmol h?1 g?1, K s sur: 476.5 μM and nitrate V max sur: 10.7 μmol h?1 g?1, K s sur: 6.1 μM) and during the assimilation phase (ammonium: V max ass: 135.6 μmol h?1 g?1, K s ass: 407.2 μM and nitrate V max ass: 1.9 μmol h?1 g?1, K s ass: 2.2 μM), with ammonium rates always higher than those of nitrate. Even though a net phosphate disappearance was observed in all treatments, uptake kinetics of this ion could not be properly estimated by the employed methodology.  相似文献   

11.
Four new platinum(II) complexes: PtII L1·H2O (C1, H2 L1 = C20H16N2O2), PtII L2Cl2 (C2, L2 = C22H16N2O2), PtII L3Cl2·H2O (C3, L3 = C20H16N2), PtII L4Cl2·0.4H2O (C4, L4 = C18H14N4) have been synthesized and characterized by using various physico-chemical techniques. The binding interaction of the four platinum(II) complexes C1C4 with calf thymus (CT)-DNA has been investigated by UV–Vis and fluorescence emission spectrometry. The apparent binding constant (K app) values follow the order: C3 > C1 > C2 > C4. In addition, fluorescence spectrometry of bovine serum albumin (BSA) with the four platinum(II) complexes C1C4 showed that the quenching mechanism might be a static quenching procedure. For C1C4, the number of binding sites was about one for BSA and the binding constants follow the order: C3 (7.08 × 105M?1) > C1 (2.82 × 105M?1) > C2 (0.85 × 105M?1) > C4 (0.15 × 105M?1). With the single condition change such as absence of an external agent, the DNA cleavage abilities of C3 exhibit remarkable changes. In addition, the cytotoxicity of C3 in vitro on tumor cells lines (MCF-7, HepG2 and HT29) were examined by MTT and showed better antitumor effects on the tested cells.  相似文献   

12.
Nitrate fertilization has been shown to increase Zn hyperaccumulation by Noccaea caerulescens (Prayon) (formerly Thlaspi caerulescens). However, it is unknown whether this increased hyperaccumulation is a direct result of NO3 ? nutrition or due to changes in rhizosphere pH as a result of NO3 ? uptake. This paper investigated the mechanism of NO3 ?-enhanced Zn hyperaccumulation in N. caerulescens by assessing the response of Zn uptake to N form and solution pH. Plants were grown in nutrient solution with 300 μM Zn and supplied with either (NH4)2SO4, NH4NO3 or Ca(NO3)2. The solutions were buffered at either pH 4.5 or 6.5. The Zn concentration and content were much higher in shoots of NO3 ?-fed plants than in NH4 +-fed plants at pH 4.5 and 6.5. The Zn concentration in the shoots was mainly enhanced by NO3 ?, whereas the Zn concentration in the roots was mainly enhanced by pH 6.5. Nitrate increased Zn uptake in the roots at pH 6.5 and increased apoplastic Zn at pH 4.5. Zinc and Ca co-increased and was found co-localized in leaf cells of NO3 ?-fed plants. We conclude that NO3 ? directly enhanced Zn uptake and translocation from roots to shoots in N. caerulescens.  相似文献   

13.
Summary The ability of Pseudomonas aeruginosa to accumulate Cd(II) ions from wastewater industries was experimentally investigated and mathematically modelled. From the potentiometric titration and non-ideal competitive analysis (NICA) model, it was found that the biomass contains three acidic sites. The values of proton binding (pK i =1.66±3.26×10−3, 1.92±1.63×10−4 and 2.16±3.79×10−4) and binding constant of cadmium metal ions (pK M1=1.99±2.45×10−3 and pK M2=1.67±4.08×10−3) on the whole surface of biomass showed that protonated functional groups and biosorption of Cd(II) ions could be attributed to a monodentate binding to one acidic site, mainly the carboxylic group. From the isothermal sorption experimental data and Langmuir model, it was also found that the value of Langmuir equilibrium (pK f) constant is 2.04±2.1×10−5 suggesting that the carboxyl group is the main active binding site. In addition, results showed that the maximum cadmium capacity (q max) and affinity of biomass towards cadmium metal ions (b) at pH 5.1 and 20 min were 96.5±0.06 mg/g and 3.40×10−3± 2.10×10−3, respectively. Finally, interfering metal ions such as Pb(II), Cu(II), Cr(III), Zn(II), Fe(II), Mn(II), Ca(II) and Mg(II) inhibited Cd(II) uptake. Comparing the biosorption of Cd(II) by various Pseudomonas isolates from contaminated environment samples (soil and sewage treatment plant) showed that maximum capacities and equilibrium times were different, indicating that there was a discrepancy in the chemical composition between biomasses of different strains.  相似文献   

14.
A copper(II) complex containing aspartame (APM) as ligand, Cu(APM)2Cl2·2H2O, was synthesized and characterized. In vitro binding interaction of this complex with human serum albumin (HSA) was studied at physiological pH. Binding studies of this complex with HSA are useful for understanding the Cu(APM)2Cl2·2H2O–HSA interaction mechanism and providing guidance for the application and design of new and more efficient artificial sweeteners drive. The interaction was investigated by spectrophotometric, spectrofluorometric, competition experiment and circular dichroism. Hyperchromicity observed in UV absorption band of Cu(APM)2Cl2·2H2O. A strong fluorescence quenching reaction of HSA to Cu(APM)2Cl2·2H2O was observed and the binding constant (Kf) and corresponding numbers of binding sites (n) were calculated at different temperatures. Thermodynamic parameters, enthalpy change (?H) and entropy change (?S) were calculated to be ?458.67 kJ mol?1 and ?1,339 J mol?1 K?1 respectively. According to the van’t Hoff equation, the reaction is predominantly enthalpically driven. In conformity with experimental results, we suggest that Cu(APM)2Cl2·2H2O interacts with HSA. In comparison with previous study, it is found that the Cu(II) complex binds stronger than aspartame.  相似文献   

15.
The potential of nonliving biomass of Hydrilla verticillata to adsorb Pb(II) from an aqueous solution containing very low concentrations of Pb(II) was determined in this study. Effects of shaking time, contact time, biosorbent dosage, pH of the medium, and initial Pb(II) concentration on metal-biosorbent interactions were studied through batch adsorption experiments. Maximum Pb(II) removal was obtained after 2 h of shaking. Adsorption capacity at the equilibrium increased with increasing initial Pb(II) concentration, whereas it decreased with increasing biosorbent dosage. The optimum pH of the biosorption was 4.0. Surface titrations showed that the surface of the biosorbent was positively charged at low pH and negatively charged at pH higher than 3.6. Fourier transform infrared (FT-IR) spectra of the biosorbent confirmed the involvement of hydroxyl and C?O of acylamide functional groups on the biosorbent surface in the Pb(II) binding process. Kinetic and equilibrium data showed that the adsorption process followed the pseudo-second-order kinetic model and both Langmuir and Freundlich isothermal models. The mean adsorption energy showed that the adsorption of Pb(II) was physical in nature. The monolayer adsorption capacity of Pb(II) was 125 mg g?1. The desorption of Pb(II) from the biosorbent by selected desorbing solutions were HNO3 > Na2CO3 > NaOH > NaNO3.  相似文献   

16.
Two cyclic and branched peptides (PLA and AZU) were synthesized with the aim of reproducing the active site of the blue copper proteins plastocyanin and azurin. Both peptides, designed on the basis of the x-ray structures of Poplar plastocyanin and Alcaligenes denitrificans azurin. contain the same coordinating residues of the parent native proteins. The visible spectra of PLA in the presence of equimolar amount of Cu(II) strongly support the interaction between the peptide and copper(II) ion. The CD titration of AZU with the Hg(II) ion indicates for the formation of two species. [A ZUHg]+ and [A ZUHg2]3+ having binding constants (Keq) of 3.106 and 2–104M?1, respectively. © 1994 John Wiley & Sons, Inc.  相似文献   

17.
The interaction between K2Cr2O7 and urease was investigated using fluorescence, UV-vis absorption, and circular dichroism (CD) spectroscopy. The experimental results showed that the fluorescence quenching of urease by K2Cr2O7 was a result of the formation of K2Cr2O7–urease complex. The apparent binding constant K A between K2Cr2O7 and urease at 295, 302, and 309 K were obtained to be 2.14?×?104, 1.96?×?104, and 1.92?×?104 L mol?1, respectively. The thermodynamic parameters, Δ and Δ were estimated to be ?5.90 kJ mol?1, 43.67 J mol?1 K?1 according to the Van’t Hoff equation. The electrostatic interaction played a major role in stabilizing the complex. The distance r between donor (urease) and acceptor (K2Cr2O7) was 5.08 nm. The effect of K2Cr2O7 on the conformation of urease was analyzed using UV-vis absorption, CD, synchronous fluorescence spectroscopy, and three-dimensional fluorescence spectra, the environment around Trp and Tyr residues were altered.  相似文献   

18.
Four complexes [Pd(L)(bipy)Cl]·4H2O (1), [Pd(L)(phen)Cl]·4H2O (2), [Pt(L)(bipy)Cl]·4H2O (3), and [Pt(L)(phen)Cl]·4H2O (4), where L = quinolinic acid, bipy = 2,2’-bipyridyl, and phen = 1,10-phenanthroline, have been synthesized and characterized using IR, 1H NMR, elemental analysis, and single-crystal X-ray diffractometry. The binding of the complexes to FS-DNA was investigated by electronic absorption titration and fluorescence spectroscopy. The results indicate that the complexes bind to FS-DNA in an intercalative mode and the intrinsic binding constants K of the title complexes with FS-DNA are about 3.5?×?104 M?1, 3.9?×?104 M?1, 6.1?×?104 M?1, and 1.4?×?105 M?1, respectively. Also, the four complexes bind to DNA with different binding affinities, in descending order: complex 4, complex 3, complex 2, complex 1. Gel electrophoresis assay demonstrated the ability of the Pt(II) complexes to cleave pBR322 plasmid DNA.  相似文献   

19.
The interaction of a novel macrocyclic copper(II) complex, ([CuL(ClO4)2] that L is 1,3,6,10,12,15-hexaazatricyclo[13.3.1.16,10]eicosane) with calf thymus DNA (ct-DNA) was investigated by various physicochemical techniques and molecular docking at simulated physiological conditions (pH = 7.4). The absorption spectra of the Cu(II) complex with ct-DNA showed a marked hyperchroism with 10 nm blue shift. The intrinsic binding constant (Kb) was determined as 1.25 × 104 M?1, which is more in keeping with the groove binding with DNA. Furthermore, competitive fluorimetric studies with Hoechst33258 have shown that Cu(II) complex exhibits the ability to displace the ct-DNA-bound Hoechst33258 indicating that it binds to ct-DNA in strong competition with Hoechst33258 for the groove binding. Also, no change in the relative viscosity of ct-DNA and fluorescence intensity of ct-DNA-MB complex in the present of Cu(II) complex is another evidence to groove binding. The thermodynamic parameters are calculated by van't Hoff equation, which demonstrated that hydrogen bonds and van der Waals interactions played major roles in the binding reaction. The experimental results were in agreement with the results obtained via molecular docking study.  相似文献   

20.
The accumulation of arginine in the cerebrospinal fluid and brains of patients suffering from acute neurodegenerative diseases like Alzheimer’s disease, point to defects in the metabolic pathways involving this amino acids. The deposits of neurofibrillary tangles and senile plaques perhaps as a consequence of fibrillogenesis of β-amyloid peptides has also been shown to be a hallmark in the aetiology of certain neurodegenerative diseases. Peptidylarginine deiminase (PAD II) is an enzyme that uses arginine as a substrate and we now show that PAD II not only binds with the peptides Aβ1-40, Aβ22-35, Aβ17-28, Aβ25-35 and Aβ32-35 but assists in the proteolytic degradation of these peptides with the concomitant formation of insoluble fibrils. PAD was purified in 12.5% yield and 137 fold with a specific activity of 59 μmol min?1?mg?1 from bovine brain by chromatography on diethylaminoethyl (DEAE)-Sephacel. Characterisation of the enzyme gave a pH and temperature optima of 7.5°C and 68°C, respectively, and the enzyme lost 50% activity within 38 min at this temperature. Michaelis-Menten kinetics established a V max and K m of 1.57 μmol min?1?ml?1 and 1.35 mM, respectively, with N-benzoyl arginine ethyl ester as substrate. Kinetic analysis was used to measure the affinity (K i) of the amyloid peptides to PAD with values between 1.4 and 4.6 μM. The formation of Aβ fibrils was rate limiting involving an initial lag time of about 24 h that was dependent on the concentration of the amyloid peptides. Turbidity measurements at 400 nm, Congo Red assay and Thioflavin-T staining fluorescence were used to establish the aggregation kinetics of PAD-induced fibril formation.  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