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.
Although metal ions can promote amyloid formation from many proteins, their effects on the formation of amyloid from transthyretin have not been previously studied. We therefore screened the effects of Cu(II), Zn(II), Al(III), and Fe(III) on amyloid formation from wild-type (WT) transthyretin as well as its V30M, L55P, and T119M mutants. Cu(II) and Zn(II) promoted amyloid formation from the L55P mutant of transthyretin at pH 6.5 but had little effect on amyloid formation from the other forms of the protein. Zn(II) promoted L55P amyloid formation at pH 7.4 but Cu(II) inhibited it. Cu(II) gave dose-dependent quenching of the tryptophan fluorescence of transthyretin and the fluorescence of 1-anilino-8-naphthalene sulfonate bound to it. Zn(II) gave dose-dependent quenching of the tryptophan but not the 1-anilino-8-naphthalene sulfonate fluorescence. Apparent dissociation constants for Cu(II) and Zn(II) binding at pH 7.4 of approximately 10 nM and approximately 1 microM (approximately 0.4 microM and approximately 5 microM at pH 6.5), respectively, were obtained from the quenching data. Zn(II) enhanced urea-mediated the dissociation of the L55P but not the WT transthyretin tetramer. Cu(II), depending on its concentration, either had no effect or stabilized the WT tetramer but could enhance urea-mediated dissociation of L55P.  相似文献   

2.
Cyclam‐based ligands and their complexes are known to show antitumor activity. This study was undertaken to examine the interaction of a diazacyclam‐based macrocyclic copper(II) complex with bovine serum albumin (BSA) under physiological conditions. The interactions of different metal‐based drugs with blood proteins, especially those with serum albumin, may affect the concentration and deactivation of metal drugs, and thereby influence their availability and toxicity during chemotherapy. In this vein, several spectral methods including UV–vis absorption, fluorescence and circular dichroism (CD) spectroscopy techniques were used. Spectroscopic analysis of the fluorescence quenching confirmed that the Cu(II) complex quenched BSA fluorescence intensity by a dynamic mechanism. In order to further determine the quenching mechanism, an analysis of Stern–Volmer plots at various concentrations of BSA was carried out. It was found that the KSV value increased with the BSA concentration. It was suggested that the fluorescence quenching process was a dynamic quenching rather than a static quenching mechanism. Based on Förster's theory, the average binding distance between the Cu(II) complex and BSA (r) was found to be 4.98 nm; as the binding distance was less than 8 nm, energy transfer from BSA to the Cu(II) complex had a high possibility of occurrence. Thermodynamic parameters (positive ΔH and ΔS values) and measurement of competitive fluorescence with 1‐anilinonaphthalene‐8‐sulphonic acid (1,8‐ANS) indicated that hydrophobic interaction plays a major role in the Cu(II) complex interaction with BSA. A Job's plot of the results confirmed that there was one binding site in BSA for the Cu(II) complex (1:1 stoichiometry). The site marker competitive experiment confirmed that the Cu(II) complex was located in site I (subdomain IIA) of BSA. Finally, CD data indicated that interaction of the Cu(II) complex with BSA caused a small increase in the α‐helical content. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

3.
We analysed the theory of the coupled equilibria between a metal ion, a metal ion-binding dye and a metal ion-binding protein in order to develop a procedure for estimating the apparent affinity constant of a metal ion:protein complex. This can be done by analysing from measurements of the change in the concentration of the metal ion:dye complex with variation in the concentration of either the metal ion or the protein. Using experimentally determined values for the affinity constant of Cu(II) for the dye, 2-(5-bromo-2-pyridylaxo)-5-(N-propyl-N-sulfopropylamino) aniline (5-Br-PSAA), this procedure was used to estimate the apparent affinity constants for formation of Cu(II):transthyretin, yielding values which were in agreement with literature values. An apparent affinity constant for Cu(II) binding to α-synuclein of ∼1 × 109 M−1 was obtained from measurements of tyrosine fluorescence quenching by Cu(II). This value was in good agreement with that obtained using 5-Br-PSAA. Our analysis and data therefore show that measurement of changes in the equilibria between Cu(II) and 5-Br-PSAA by Cu(II)-binding proteins provides a general procedure for estimating the affinities of proteins for Cu(II).  相似文献   

4.
Ghatak  H.  Mukhopadhyay  S.K.  Jana  T.K.  Sen  B.K.  Sen  S. 《Wetlands Ecology and Management》2004,12(3):145-155
Humic (HA) and fulvic (FA) acids isolated from mangrove sediments of Sundarban, the largest delta on earth in the estuarine phase of the river Ganges, were studied and attempts were made to characterize their binding sites by quenching of Synchronous fluorescence (SyF) bands with Fe (III) and Cu (II). A modified Stern-Volmer relationship applicable for static quenching was applied for the determination of conditional stability constants and the data were compared with those determined by potentiometric titration. In the excited state HA and FA showed different acidity constant compared to the ground state. Values of the conditional stability constant (log Kc) for Fe (III) and Cu (II) indicated that binding sites were bidentate in nature. FA were better chelators than the HA fractions. High energy binding sites of both FA & HA were occupied by Fe(III) and the low energy binding sites, mainly responsible for mobilization and immobilization of metal, were occupied by Cu(II).  相似文献   

5.
Sorption of Cu(II) and Cd(II) onto the extracellular polymeric substances (EPS) produced by Aspergillus fumigatus was investigated for the initial pH of the solution, EPS concentrations, contact time, NaCl concentration, initial metal ion concentration and the presence of other ions in the solution. The results showed that the adsorption of metal ions was significantly affected by pH, EPS concentrations, initial metal concentration, NaCl concentration and co-ions. The sorption of Cu(II) and Cd(II) increased with increasing pH and initial metal ion concentration but decreased with an increase in the NaCl concentration. The maximum sorption capacities of A. fumigatus EPS calculated from the Langmuir model were 40 mg g−1 EPS and 85.5 mg g−1 EPS for Cu(II) and Cd(II), respectively. The binary metal sorption experiments showed a selective metal binding affinity in the order of Cu(II) > Pb(II) > Cd(II). Both the Freundlich and Langmuir adsorption models described the sorption of Cu(II) and Cd(II) by the EPS of Afumigatus adequately. Fourier transform infrared spectroscopy (FTIR) analysis revealed that carboxyl, amide and hydroxyl functional groups were mainly correlated with the sorption of Cu(II) and Cd(II). Energy dispersive X-ray (EDX) system analysis revealed that the ion-exchange was an important mechanism involved in the Cu(II) and Cd(II) sorption process taking place on EPS.  相似文献   

6.
An aza-oxa-thia macrocycle, 5,14-dioxa-2,17-dithia[6](1,2)benzeno[6](2,6)pyridinophane, L1, the related smaller macrocycle 2,14-dithia-11-oxa-[3](1,2)benzeno[6](2,6)pyridinophane, L2, and the complexes with Pd(II) and Cu(II) of the macrocycle, L1, have been synthesized. The crystal structure of L2 and those of the two metal complexes have been determined. In the complexes, the metal ions adopt exclusively square planar geometry in which the pyridine nitrogen, two sulfurs and one chlorine atom are coordinated and there is no appreciable interaction with the oxygen donors. Thus, the `hard-soft acid-base' principle is illustrated by the behaviour of L1. The structures of both complexes are compared with the previously reported mixed aza-thia macrocycle, 2,5,14,17-tetrathia[6](1,2)benzeno[6](2,6)pyridinophane. The crystal structure of the smaller macrocycle, L2, is also discussed and due to the nature of its smaller cavity, attempts to make complexes with it have not been successful.  相似文献   

7.
Cu(I) and Cu(II) form stable 1:1 complexes with bleomycin (BLM). The affinity of both metals for the drug is greater than that of Fe(II). Cu(I) . BLM A2 binds to calf thymus DNA with about the same affinity as Fe(II) . BLM, as judged by DNA-induced fluorescence quenching of the bithiazole moiety of BLM. Based on 1H NMR and potentiometric titration data, the Cu(I) complexes of BLM are shown to have geometries very different than those of other BLM . metal(II) complexes studied thus far. As Cu(I) . BLM is oxidation-reduction active, its geometry is of importance in defining the structural requirements for BLM activity.  相似文献   

8.
Copper (Cu(II)) and nickel (Ni(II)) are often encountered in wastewaters. This study investigated the individual toxic effects of long-term addition of Cu(II) and Ni(II) on the biochemical properties of aerobic granules in sequencing batch reactors (SBRs). The biochemical properties of aerobic granules were characterized by extracellular polymeric substances (EPS) content, dehydrogenase activity, microbial community biodiversity, and SBR performance. One SBR was used as a control system, while another two received respective concentration of Cu(II) and Ni(II) equal to 5 mg/L initially and increased to 15 mg/L on day 27. Results showed that the addition of Cu(II) drastically reduced the biomass concentration, bioactivity, and biodiversity of aerobic granules, and certainly deteriorated the treatment performance. The toxic effect of Ni(II) on the biodiversity of aerobic granules was milder and the aerobic granular system elevated the level of Ni(II) toxicity tolerance. Even at a concentration of 15 mg/L, Ni(II) still stimulated the biomass yield and bioactivity of aerobic granules to some extent. The elevated tolerance seemed to be owed to the concentration gradient developed within granules, increased biomass concentration, and promoted EPS production in aerobic granular systems.  相似文献   

9.
The interaction of Cu(II) with human lactoferrin has been studied as a function of pH, using electronic and electron spin resonance spectroscopy. Specific Cu(II) binding, with bicarbonate as the co-anion, occurs over the pH range 6 to 9. In the presence of a fiftyfold molar excess of oxalate, a monocopper(II) lactoferrin oxalate complex forms when the Cu(II) to protein is 1:1. If this ratio is increased to 2:1, a hybrid complex forms, in which the second copper utilizes bicarbonate as the co-anion, thus demonstrating, as for serum transferrin, a difference in the anion binding sites. The quenching of the intrinsic fluorescence of apolactoferrin is significantly less in the presence of oxalate than bicarbonate. The interaction of Cu(II) with apolactoferrin in the presence of the malonate, glycolate, thioglycolate, glycinate, and ethylenediaminetetraacetate ions has been examined.  相似文献   

10.
This paper reports biosorption of Zn(II), Cu(II) and Co(II) onto O. angustissima biomass from single, binary and ternary metal solutions, as a function of pH and metal concentrations via Central Composite Design generated by statistical software package Design Expert 6.0. The experimental design revealed that metal interactions could be best studied at lower pH range i.e. 4.0-5.0, which facilitates adequate availability of all the metal ions. The sorption capacities for single metal decreased in the order Zn(II)>Co(II)>Cu(II). In absence of any interfering metals, at pH 4.0 and an initial metal concentration of 0.5 mM in the solution, the adsorption capacities were 0.33 mmol/g Zn(II), 0.26 mmol/g Co(II) and 0.12 mmol/g Cu(II). In a binary system, copper inhibited both Zn(II) and Co(II) sorption but the extent of inhibition of former was greater than the latter; sorption values being 0.14 mmol/g Zn(II) and 0.27 mmol/g Co(II) at initial Zn(II) and Co(II) concentration of 1.5 mM each, pH 4.0 and 1mM Cu(II) as the interfering metal. Zn(II) and Co(II) were equally antagonistic to each others sorption; Zn(II) and Co(II) sorption being 0.23 and 0.24 mmol/g, respectively, at initial metal concentration of 1.5 mM each, pH 4.0 and 1mM interfering metal concentration. In contrast, Cu(II) sorption remained almost unaffected at lower concentrations of the competing metals. Thus, in binary system inhibition dominance observed was Cu(II)>Zn(II), Cu(II)>Co(II) and Zn(II) approximately Co(II), due to this the biosorbent exhibited net preference/affinity for Cu(II) sorption over Zn(II) or Co(II). Hence, the affinity series showed a trend of Cu(II)>Co(II)>Zn(II). In a ternary system, increasing Co(II) concentration exhibited protection against the inhibitory effect of Cu(II) on Zn(II) sorption. On the other hand, the inhibitory effect of Zn(II) and Cu(II) on Co(II) sorption was additive. The model equation for metal interactions was found to be valid within the design space.  相似文献   

11.
By using an LKB2277 Bioactivity Monitor, stop-flow mode, the power-time curves of Escherichia coli at 37 degrees C affected by Cu(II) were determined. Some parameters, such as growth rate constants k, inhibitory ratio I, the heat output Qlog in the log phase, and the generation times G were obtained. According to these parameters, we found that a low concentration of Cu(II) (0-20 microg/mL) had an promoting action on the growth of E. coli, but a high concentration of Cu(II) (40-100 microg/mL) had an inhibitory action. The toxicity of Cu(II) can also be expressed as the half-inhibitory concentration IC50; the value is 69.7 microg/mL. The assay is quantitative, inexpensive, and versatile.  相似文献   

12.
Shields SB  Franklin SJ 《Biochemistry》2004,43(51):16086-16091
A chimeric Cu-binding peptide has been designed on the basis of a turn substitution of the prion (PrP) octarepeat Cu-binding site into the engrailed homeodomain helix-turn-helix motif (HTH). This system is a model for the investigation of a single PrP Cu-binding site in a defined protein context. The 28-mer Cu-HTH peptide P7 spectroscopically mimics the PrP octarepeat (P7 = TERRRQQLSHGGGWGEAQIKIWFQNKRA). The Cu(II)-binding affinity of P7 was determined by ESI-MS and tryptophan fluorescence titrations to be K(d) = 2.5 +/- 0.7 microM at pH = 7.0. The quenching of fluorescence of the Trp within the binding loop (underlined above) is pH dependent and highly specific for Cu(II). No Trp quenching was observed in the presence of divalent Zn, Mn, Co, Ni, or Ca ions, and ESI-MS titrations confirmed that these divalent ions do not appreciably bind to P7. The EPR spectrum of Cu(II)-P7 shows that the Cu environment is axial and consistent with 6-coordinate N(3)O(H(2)O)(2) or N(4)(H(2)O)(2) coordination (A( parallel) = 172 x10(-)(4) cm(-)(1); g( parallel) = 2.27), very similar to that of the PrP octarepeat itself. Also like PrP, circular dichroism studies show that apo P7 is predominantly disordered in solution, and the structure is slightly enhanced by Cu binding. These data show the Cu-PrP HTH peptide reproduces the Cu-binding behavior of a single PrP octarepeat in a new context.  相似文献   

13.
In the presence of a nonlethal concentration of Cu(II), washed Escherichia coli ATCC11775 cells were killed by (-)-epigallocatechin (EGC) and (-)-epicatechin (EC). Cell killing was accompanied by a depletion in both the ATP and potassium pools of the cells, but the DNA double strand was not broken, indicating that the bactericidal activity of catechins in the presence of Cu(II) results from damage to the cytoplasmic membrane. Induction of endogenous catalase in E. coli cells increased their resistance to being killed by the combination of catechins and Cu(II). In all cases studied, EGC and EC with Cu(II) were found to generate hydrogen peroxide, but its concentration was too low to account for the bactericidal activity. The bactericidal activity of EGC in the presence of Cu(II) was completely suppressed by ethylenediaminetetraacetate, bathocuproine, catalase, superoxide disumutase (SOD), heated catalase, and heated SOD, but not by dimethyl sulfoxide. When catalase, either heated or unheated, was added to the cells incubated with EGC in the presence of Cu(II), it completely inhibited further killing of the cells. These findings suggest that recycling redox reactions between Cu(II) and Cu(I), involving catechins and hydrogen peroxide on the cell surface, must be important in the mechanism of the killing.  相似文献   

14.
Lu Y  Li XR  He MZ  Su YG  Zeng FJ 《应用生态学报》2011,22(4):936-942
以骆驼蓬幼苗为材料,采用盆栽试验研究不同浓度(0、50、100、200、400 mg·kg-1)Ni、Cu处理对骆驼蓬叶片光合作用、叶绿素荧光特性及生长状况的影响.结果表明:随着Ni浓度的增加,骆驼蓬幼苗叶片的光合色素含量、净光合速率(Pn)、气孔导度(Gs)、蒸腾速率(Tr)、PS Ⅱ最大光化学效率(Fv/Fm)、PS Ⅱ电子传递量子产率(φpsⅡ)、光化学猝灭系数(qp)及各项生长指标均呈显著下降趋势,而细胞间隙CO2浓度(Ci)和非光化学猝灭系数(qn)呈显著增加趋势,其中Pn的下降主要是由非气孔限制所致;骆驼蓬幼苗叶片的光合色素含量、Pn、Gs、Tr、Ci、Fv/Fm、φpsⅡ、qp及各项生长指标均在50 mg·kg-1Cu处理时达到峰值,叶绿素a和b、Pn、Gs、Tr、Ci、Fv/Fm及各项生长指标值在100 mg·kg-1Cu处理时仍微高于对照,而后随Cu浓度的增加,光合色素含量、Pn、Gs、Tr、Ci、Fv/Fm、φpsⅡ、qp及各项生长指标均呈下降趋势,qN呈增加趋势,其中Pn的下降主要是由气孔限制所致.  相似文献   

15.
The crystal and molecular structures of the N-rac-isomer of the nickel(II) complex of 14-membered amide-containing macrocycle [NiL1] · 4H2O (H2L1=5,12-dioxo-1,4,8,11-tetraazacyclotetradecane) have been determined. Two deprotonated amide and two amine donors co-ordinate to the nickel(II) in nearly square planar manner with Ni-Namine bonds longer than Ni-Namide ones (1.930 vs. 1.898 Å). Water molecules do not co-ordinate and form hydrogen bond bridges between macrocyclic units in the crystal lattice. The analysis of 1H NMR data confirmed that the solid-state conformation of the macrocycle in N-rac[NiL1] is retained in aqueous solution though equilibrated with some amount of N-meso isomer. The comparison of the spectroscopic characteristics of the M(II) and M(III) complexes and the redox potentials of M(III/II) couples (M=Ni and Cu) for ML1 with those for ML2(H2L2=5,7-dioxo-1,4,8,11-tetraazacyclotetradecane) revealed a rather small influence of the trans- vs. cis-arrangement of amide donors in co-ordination spheres of the metal ions.  相似文献   

16.
The kinetics of reconstitution of Pseudomonas aeruginosa azurin from its apoprotein and copper(II) salts have been studied using absorbance at 625 nm and fluorescence emission at 308 nm as monitors of the process. At low Cu(II) concentrations the rates of both absorbance and fluorescence changes are linearly dependent on Cu(II) concentration. At higher Cu(II) concentrations the rate of absorbance change is independent of Cu(II) concentration. The rates of both absorbance and fluorescence changes as a function of pH suggest that the titration of a single ionizable group is important for the Cu(II)-dependent reaction. Overall analysis of the kinetics suggests that the fluorescence change and the absorbance change are associated with at least two steps in the overall pathway of the formation of the metal-protein complex, and that the copper(II) and tryptophan environments in this protein, though perhaps spatially close, may be distinct.  相似文献   

17.
The biosorption from artificial wastewaters of heavy metals (Cd(II), Pb(II) and Cu(II)) onto the dry fungal biomass of Phanerochaete chryosporium was studied in the concentration range of 5-500 mg l(-1). The maximum absorption of different heavy metal ions on the fungal biomass was obtained at pH 6.0 and the biosorption equilibrium was established after about 6 h. The experimental biosorption data for Cd(II), Pb(II) and Cu(II) ions were in good agreement with those calculated by the Langmuir model.  相似文献   

18.
Srikanth R  Wilson J  Burns CS  Vachet RW 《Biochemistry》2008,47(35):9258-9268
While the Cu(II) binding sites of the prion protein have been well studied under Cu-saturation conditions, the identity of the residues involved in coordinating Cu(II) at low stoichiometries and the order in which the binding sites load with Cu(II) remain unresolved. In this study, we have used two mass spectrometry based methods to gather insight into Cu(II)-prion binding under different stoichiometric loadings of Cu(II). The first method uses metal-catalyzed oxidation reactions to site specifically modify the residues bound to Cu(II) in solution, and the second method determines Cu binding sites based on the protection of His from modification by diethyl pyrocarbonate when this residue binds Cu(II) in solution. For both methods, the residues that are labeled by these reactions can then be unambiguously identified using tandem mass spectrometry. Upon applying these two complementary methods to a construct of the prion protein that contains residues 23-28 and 57-98, several noteworthy observations are made. Coordination of Cu(II) by multiple His imidazoles is found at 1:1 and 1:2 PrP:Cu(II) ratios. Notably, there appear to be four to seven isomers of this multiple histidine coordination mode in the 1:1 complex. Furthermore, our data clearly show that His96 is the dominant Cu(II) binding ligand, as in every isomer His96 is bound to Cu(II). The individual octarepeat binding sites begin to fill at ratios of 1:3 PrP:Cu(II) with no clear preference for the order in which they load with Cu(II), although the His77 octarepeat appears to saturate last. The existence of several "degenerate" Cu binding modes at low PrP:Cu(II) ratios may allow it to more readily accept additional Cu(II) ions, thus allowing PrP to transition from a singly Cu(II) bound state to a multiply Cu(II) bound state as a function of cellular Cu(II) concentration.  相似文献   

19.
The true Michaelis constant for GSH and CDNB was 0.287 mM and 0.180 mM, respectively. Regarding the quantitative effect of Cu(II) and Cd(II) inhibition on the GST system, the I50 value for Cu(II) was 0.250 mM; in contrast, Cd(II) GST-inhibition did not reach the I50 value. When the varied substrate was GSH and CDNB was fixed at saturant concentration, the Cu(II)-inhibition was consistent with a pure competitive pattern. However a mixed pattern was found when CDNB was the varied substrate and GSSH was fixed at saturant concentration. The Cd(II) inhibition effect was consistent with an uncompetitive pattern when GSH was the varied substrate and CDNB was kept at saturant level. When CDNB changed over an extensive range of concentration, the inhibition effect shows a mixed inhibition pattern with a competitive character. In addition the inhibition constants of Cu(II) were one order of magnitude lower than those of Cd(II).  相似文献   

20.
Poly(N-vinylimidazole), PVIm, gels were prepared by γ-irradiation polymerization of N-vinylimidazole in aqueous solutions. These affinity gels with a water swelling ratio of 1800% for plain polymeric gel and between 30 and 80% for Cu(II) and Co(II)-chelated gels at pH 6.0 in phosphate buffer were used in glucose oxidase (GOx) adsorption–desorption studies. Different amounts of Cu(II) and Co(II) ions (maximum 3.64 mmol/g dry gel for Cu(II) and 1.72 mmol/g dry gel for Co(II)) were loaded onto the gels by changing the initial concentration of Cu(II) and Co(II) ions, and pH. GOx adsorption on these gels from aqueous solutions containing different amount of GOx at different pH was investigated in batch reactors. Immobilized glucose oxidase activity onto the poly(N-vinylimidazole), and Cu(II) and Co(II)-chelated poly(N-vinylimidazole) were investigated with changing pH and the initial glucose oxidase concentration. Maximum activity of immobilized glucose oxidase onto the PVIm, Cu(II) and Co(II)-chelated PVIm gels was investigated and pH dependence was observed to be at pH 6.5 for free enzyme, pH 7.0 for PVIm, pH 7.5 for Cu(II) and Co(II)-chelated PVIm gels, respectively. The stability of the immobilized enzyme is very high for all gels and the residual activity was higher than 93% in the first 10 days.  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