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 671 毫秒
1.
Karr JW  Szalai VA 《Biochemistry》2008,47(17):5006-5016
Copper has been proposed to play a role in Alzheimer's disease through interactions with the amyoid-beta (Abeta) peptide. The coordination environment of bound copper as a function of Cu:Abeta stoichiometry and Abeta oligomerization state are particularly contentious. Using low-temperature electron paramagnetic resonance (EPR) spectroscopy, we spectroscopically distinguish two Cu(II) binding sites on both soluble and fibrillar Abeta (for site 1, A parallel = 168 +/- 1 G and g parallel = 2.268; for site 2, A parallel = 157 +/- 2 G and g parallel = 2.303). When fibrils that have been incubated with more than 1 equiv of Cu(II) are washed, the second Cu(II) ion is removed, indicating that it is only weakly bound to the fibrils. No change in the Cu(II) coordination environment is detected by EPR spectroscopy of Cu(II) with Abeta (1:1 ratio) collected as a function of Abeta fibrillization time, which indicates that the Cu(II) environment is independent of Abeta oligomeric state. The initial Cu(II)-Abeta complexes go on to form Cu(II)-containing Abeta fibrils. Transmission electron microscopy images of Abeta fibrils before and after Cu(II) addition are the same, showing that once incorporated, Cu(II) does not affect fibrillar structure; however, the presence of Cu(II) appears to induce fibril-fibril association. On the basis of our results, we propose a model for Cu(II) binding to Abeta during fibrillization that is independent of peptide oligomeric state.  相似文献   

2.
Structural changes associated with the exposure of human serum albumin (HSA) to glucose with or without the presence of Cu (II) have been characterized using a bank of methods for structural analysis including circular dichroism (CD), amino acid analysis (AAA), fluorescence measurements, SDS-PAGE, and boronate binding (which is a measure of Amadori product formation). We show that in the short-term (10 d) incubation mixtures, HSA is resistant to Cu (II)-mediated oxidative damage and that the early products of glycation of HSA had minimal effects on the folded structure. Amino acid analysis showed that there was no formation of advanced glycation endproducts (AGE), which can be measured by loss of lysine. This remained the case in longer term incubation of HSA (56 d) in the hyperglycemic concentration range (5–25 mM glucose) despite increased levels of Amadori product (60% boronate binding) and the formation of glycophore (Excitation 350, Emission 425). At high, nonphysiological concentrations (100 mM and 500 mM) of glucose, glycophore formation increased and 3 and 11 mol Lysine-glucose adduct/mol HSA were converted to AGE, respectively. This was accompanied by increased damage to tryptophan and protein-protein crosslinking but only minor tertiary structural change. In the presence of Cu (II), however, AGE formation was accompanied by extensive damage to histidine and tryptophan side chains, main chain fragmentation, and loss of both secondary and tertiary structure. Thus, changes in structure appear to be the result of oxidation as opposed to glycation, per se. © 1997 Elsevier Science Inc.  相似文献   

3.
The copper(II) complex of 3,5-diisopropylsalicylate is a lipophilic water-insoluble binuclear complex, Cu(II) (3,5-DIPS) , that has attracted interest because of a wide range of pharmacological activities. This study was undertaken to examine bonding interactions between the complex and human serum albumin (HSA) to help elucidate the mode of transport of the complex in vivo. Electron paramagnetic resonance, numerical magnetic resonance and UV-visible absorption spectroscopic studies were performed using 200 M aqueous solutions (pH 7.5) of HSA to which had been added up to three molar equivalents of CuCl , CuSO , or Cu(II) (3,5-DIPS). Both EPR and UV-visible spectra demonstrated the presence of more than one copper bonding site on HSA, and proton NMR spectra showed that the 3,5-DIPS ligand is also bonded to HSA. These results indicate that there is no observable direct coordination of the ligand to copper in the presence of HSA, and that the majority of the copper and 3,5-DIPS bond to HSA at separate sites. Addition of solid Cu(II) (3,5-DIPS) to HSA at pH 7.5 similarly resulted in spectra that suggest that there are no ternary Cu(II)(3,5-DIPS), Cu(II)(3,5-DIPS) , or Cu(II) (3,5-DIPS) complexes formed with HSA. It is concluded that any ternary complexes formed in the presence of HSA are below the spectroscopic detection limits and represent less than 5% of the total copper. © Rapid Science 1998.  相似文献   

4.
The heat induced aggregation of human serum albumin (HSA) with and without an equimolar amount of Cu(II) and Zn(II) was investigated by using optical absorption, fluorescence, AFM and EPR spectroscopy. Turbidity experiments as a function of temperature indicate that the protein aggregation occurs after the melting of the protein. The kinetic of HSA aggregation, investigated between 60 and 70 °C by monitoring the optical density changes at 400 nm on a 180 min time window, shows an exponential growth with a rate that increases with the temperature. Fluorescence of the thioflavin T evidences a significant increase of the intensity at 480 nm at increasing incubation time. These results combined with AFM experiments show that the protein aggregates are elongated oligomers with fibrillar-like features. The absence of a lag-phase suggests that the early stage aggregation of HSA follows a downhill pathway that does not require the formation of an organized nucleus. The presence of Cu(II) and Zn(II) ions does not affect the thermally induced aggregation process and the morphology of HSA aggregates. The result is compatible with the binding of the metal ions to the protein in the native state and with the high conformational stability of HSA.  相似文献   

5.
We have studied the Cu(II), Co(II), and Fe(III) complexes of the antineoplastic drug bleomycin by using electron spin--echo envelope spectroscopy. For all three complexes, nitrogen coordination of the metal ions is demonstrated. For the Cu(II)-- and Co(II)--drug complexes, we have been able to identify imidazole as a metal ligand.  相似文献   

6.
The prolonged glycation of human serum albumin (HSA) results in significant changes in its structure. The identity of these structural changes and the influence of carbohydrates on these changes require further study. Here, we evaluated structural changes and amyloid formation of HSA upon incubation with Glc, Fru, or Rib. Fluorescence spectrophotometry, surface tension analysis, and transmission electron microscopy (TEM) were utilized to evaluate the structures of glycated HSA. The physicochemical properties including excess free energy, protein adsorption at the air-water interface, critical aggregation concentration (CAC), and surface activity indicated an increase in hydrophobicity and partial unfolding of HSA structure upon glycation. Thus, it appears that AGE products can act as detergents. Incubation of HSA with these sugars after 20 wks induced significant amyloid nanofibril formation. Together these results indicate that prolonged glycation of HSA is associated with a transition from helical structure to beta-sheet (amyloid formation).  相似文献   

7.
Copper is known to exert diverse effects on the self-association of proteins and has been found in amyloid deposits that are involved in neurodegenerative disease processes. The effects of the metal ion on the protein during fibrillation were investigated by fluorescence, circular dichroism spectroscopy and fluorescence microscopy. We report for the first time, the complete reduction of Cu(II)→Cu(I) in vitro during fibrillation of hen egg white lysozyme at pH 7. This was confirmed by the lack of any signal for Cu(II) in electron paramagnetic resonance spectroscopy and quantification of Cu(I) was achieved by a bathocuproine disulfonate based assay.  相似文献   

8.
Shin BK  Saxena S 《Biochemistry》2008,47(35):9117-9123
We provide direct evidence that all three histidine residues in amyloid-beta 1-16 (Abeta 1-16) coordinate to Cu(II). In our approach, we generate Abeta 1-16 analogues, in each of which a selected histidine residue is isotopically enriched with (15)N. Pulsed electron spin resonance (ESR) experiments such as electron spin echo envelope modulation (ESEEM) and hyperfine sublevel correlation (HYSCORE) spectroscopy clearly show that all three histidine imidazole rings at positions 6, 13 and 14 in Abeta 1-16 bind to Cu(II). The method employed here does not require either chemical side chain modification or amino acid residue replacement, each of which is traditionally used to determine whether an amino acid residue in a protein binds to a metal ion. We find that the histidine coordination in the Abeta 1-16 peptide is independent of the Cu(II)-to-peptide ratio, which is in contrast to the Abeta 1-40 peptide. The ESR results also suggest tight binding between the histidine residues and the Cu(II) ion, which is likely the reason for the high binding affinity of the Abeta peptide for Cu(II).  相似文献   

9.
The active site structure of Cu(II) in dopamine beta-hydroxylase, isolated from bovine adrenal medulla, was studied by pulsed electron paramagnetic resonance (EPR) spectroscopy. Fourier transformation of the stimulated electron spin-echo envelope revealed frequency components characteristic of Cu(II)-histidyl imidazole coordination. The three major lines in the spectrum at 0.7, 1.4, and 4.0 MHz are typical for Cu(II)-imidazole complexes where imidazole is protonated and equatorially coordinated. Quantitation of the number of imidazole ligands bound to Cu(II) in enzyme containing two, four, and eight Cu per protein tetramer, as well as characterization of the superhyperfine coupling parameters, was achieved by spectral simulation. In all cases, it was shown that there are three, or more likely four, imidazole ligands bound to Cu(II). Addition of deuteriated substrate analogues to the enzyme did not produce any observable deuterium modulation in the spin-echo envelopes, thus indicating that the distance between substrate deuterons and Cu(II) is greater than 5 A.  相似文献   

10.
The binding of Cd(II) and Zn(II) to human serum albumin (HSA) and dog serum albumin (DSA) has been studied by equilibrium dialysis and 113Cd(II)-NMR techniques at physiological pH. Scatchard analysis of the equilibrium dialysis data indicate the presence of at least two classes of binding sites for Cd(II) and Zn(II). On analysis of the high-affinity class of sites, HSA is shown to bind 2.08 +/- 0.09 (log K = 5.3 +/- 0.6) and 1.07 +/- 0.12 (log K = 6.4 +/- 0.8) moles of Cd(II) and Zn(II) per mole of protein, respectively. DSA bound 2.02 +/- 0.19 (log K = 5.1 +/- 0.8), and 1.06 +/- 0.15 (log K = 6.0 +/- 0.2) moles of Cd(II) and Zn(II) per mole of protein, respectively. Competition studies indicate the presence of one high-affinity Cd(II) site on both HSA and DSA that is not affected by Zn(II) or Cu(II), and one high-affinity Zn(II) site on both HSA and DSA that is not affected by Cd(II) or Cu(II). 113Cadmium-HSA spectra display three resonances corresponding to three different sites of complexation. In site I, Cd(II) is most probably coordinated to two or three histidyl residues, site II to one histidyl residue and three oxygen ligands (carboxylate), while for the most upfield site III, four oxygens are likely to be involved in the binding of the metal ion. The 113Cd(II)-DSA spectra display only two resonances corresponding to two different sites of complexation. The environment around Cd(II) at sites I and II on DSA is similar to sites I and II, respectively, on HSA. No additional resonances are observed in any of these experiments and in particular in the low field region where sulfur coordination occurs. Overall, our results are consistent with the proposal that the physiologically important high-affinity Zn(II) and Cd(II) binding sites of albumins are located not at the Cu(II)-specific NH2-terminal site, but at internal sites, involving mostly nitrogen and oxygen ligands and no sulphur ligand.  相似文献   

11.
Unlike human serum albumin (HSA), dog serum albumin (DSA) does not possess the characteristics of the specific first binding site for Cu(II). In DSA, the important histidine residue in the third position, responsible for the Cu(II)-binding specificity in HSA, is replaced by a tyrosine residue. In order to study the influence of the tyrosine residue in the third position of DSA, a simple model of the NH2-terminal native sequence tripeptide of DSA, glycylglycyl-L-tyrosine-N-methylamide (GGTNMA) was synthesized and its Cu(II)-binding properties studied by analytical potentiometry, spectrophotometry, CD, and NMR spectroscopy. The species analysis indicated the existence of five mono-complexes at different protonation states: MHA, MA, MH-1A, MH-2A, MH-3A, and only one bis-complex MH-2A-2. The complexing ability of GGTNMA to Cu(II) was found to be weaker than that of the Cu(II) binding peptide models of HSA. The visible absorption spectra of Cu(II)-GGTNMA complexes are similar to those observed in the case of DSA-Cu(II) complexes. The weaker binding and the spectral properties of Cu(II)-GGTNMA complexes are consistent with less specific Cu(II)-binding properties of the peptide of this sequence similar to what was noted with DSA. CD results are in excellent agreement with species analysis and visible spectra where it is clearly evident that Cu(II) binds to GGTNMA starting from the alpha-NH2 group and step by step to deprotonated amide nitrogens as the pH is raised. The absence of any charge transfer band around 400 nm strongly indicates that Cu(II) does not bind to the phenolate group. Furthermore, NMR results are consistent with the noninvolvement of the tyrosine residue of GGTNMA in Cu(II) complexation. Thus, it is clear that the low Cu(II)-binding affinity of DSA is due to the genetic substitution of tyrosine for histidine at the NH2-terminal region of the protein.  相似文献   

12.
Metal ions such as zinc and copper can have dramatic effects on the aggregation kinetics of and the structures formed by several amyloidogenic peptides/proteins. Depending on the identity of the amyloidogenic peptide/protein and the conditions, Zn(II) and Cu(II) can promote or inhibit fibril formation, and in some cases these metal ions have opposite effects. To better understand this modulation of peptide aggregation by metal ions, the impact of Zn(II) binding to three amyloidogenic peptides (Aβ14-23, Aβ11-23, and Aβ11-28) on the formation and structure of amyloid-type fibrils was investigated. Zn(II) was able to accelerate fibril formation for all three peptides as measured by thioflavin T fluorescence and transmission electron microscopy. The effects of Zn(II) on Aβ11-23 and Aβ11-28 aggregation were very different compared with the effects of Cu(II), showing that these promoting effects were metal-specific. X-ray absorption spectroscopy suggested that the Zn(II) binding to Aβ11-23 and Aβ11-28 is very different from Cu(II) binding, but that the binding is similar in the case of Aβ14-23. A model is proposed in which the different coordination chemistry of Zn(II) compared with Cu(II) explains the metal-specific effect on aggregation and the difference between peptides Aβ14-23 and Aβ11-23/Aβ11-28.  相似文献   

13.
Ha C  Ryu J  Park CB 《Biochemistry》2007,46(20):6118-6125
The abnormal deposition and aggregation of beta-amyloid (Abeta) on brain tissues are considered to be one of the characteristic neuropathological features of Alzheimer's disease (AD). Environmental conditions such as metal ions, pH, and cell membranes are associated with Abeta deposition and plaque formation. According to the amyloid cascade hypothesis of AD, the deposition of Abeta42 oligomers as diffuse plaques in vivo is an important earliest event, leading to the formation of fibrillar amyloid plaques by the further accumulation of soluble Abeta under certain environmental conditions. In order to characterize the effect of metal ions on amyloid deposition and plaque growth on a solid surface, we prepared a synthetic template by immobilizing Abeta oligomers onto a N-hydroxysuccinimide ester-activated solid surface. According to our study using ex situ atomic force microscopy (AFM), Fourier transform infrared spectroscopy (FT-IR), and thioflavin T (ThT) fluorescence spectroscopy, Cu2+ and Zn2+ ions accelerated both Abeta40 and Abeta42 deposition but resulted only in the formation of "amorphous" aggregates. In contrast, Fe3+ induced the deposition of "fibrillar" amyloid plaques at neutral pH. Under mildly acidic environments, the formation of fibrillar amyloid plaques was not induced by any metal ion tested in this work. Using secondary ion mass spectroscopy (SIMS) analysis, we found that binding Cu ions to Abeta deposits on a solid template occurred by the possible reduction of Cu ions during the interaction of Abeta with Cu2+. Our results may provide insights into the role of metal ions on the formation of fibrillar or amorphous amyloid plaques in AD.  相似文献   

14.
Transmissible spongiform encephalopathies in mammals are believed to be caused by scrapie form of prion protein (PrP(Sc)), an abnormal, oligomeric isoform of the monomeric cellular prion protein (PrP(C)). One of the proposed functions of PrP(C) in vivo is a Cu(II) binding activity. Previous studies revealed that Cu(2+) binds to the unstructured N-terminal PrP(C) segment (residues 23-120) through conserved histidine residues. Here we analyzed the Cu(II) binding properties of full-length murine PrP(C) (mPrP), of its isolated C-terminal domain mPrP(121-231) and of the N-terminal fragment mPrP(58-91) in the range of pH 3-8 with electron paramagnetic resonance spectroscopy. We find that the C-terminal domain, both in its isolated form and in the context of the full-length protein, is capable of interacting with Cu(2+). Three Cu(II) coordination types are observed for the C-terminal domain. The N-terminal segment mPrP(58-91) binds Cu(2+) only at pH values above 5.0, whereas both mPrP(121-231) and mPrP(23-231) already show identical Cu(II) coordination in the pH range 3-5. As the Cu(2+)-binding N-terminal segment 58-91 is not required for prion propagation, our results open the possibility that Cu(2+) ions bound to the C-terminal domain are involved in the replication of prions, and provide the basis for further analytical studies on the specificity of Cu(II) binding by PrP.  相似文献   

15.
The pharmacokinetics and pharmacodynamics of any drug will depend, largely, on the interaction that has with human serum albumin (HSA), the most abundant plasma protein. The interaction between newly synthesized Pd(II) complexe, 2,2'-bipyridin Butylglycinato Pd(II) nitrate, an anti-tumor component, with HSA was studied at different temperatures by fluorescence, far UV circular dichroism (CD), UV-visible spectrophotometry and theoretical approaches. The Pd(II) complex has a strong ability to quench the intrinsic fluorescence of HSA through a dynamic quenching procedure. The binding parameters and thermodynamic parameters, including δH°, δS° and δG° were calculated by fluorescence quenching method, indicated that hydrophobic forces play a major role in the interaction of Pd(II) complex with HSA. Based on Autodock, FRET (fluorescence resonance energy transfer) and fluorescence quenching data, it may be concluded that one of the binding sites in the complex of HSA is near the only one Trp of HSA (Trp214) in sub domain IIA of the protein. Far-UV-CD results indicated that Pd(II)-complex induced increase in the α-helical content of the protein. The anti-tumor property of the synthesized Pd(II) complex was studied by testing it on human tumor cell line K562. The 50% cytotoxic concentration (Cc??) of complex was determined using MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide) assay. Also, fluorescence staining with DAPI (4,6-diamidino-2-phenylindole) revealed some typical nuclear changes that are characteristic of apoptosis which is induced at Cc?? concentration of Pd(II) complex in K562 cell line after 24?h incubation. Our results suggest that Pd(II) complex is a promising anti-proliferative agent and should execute its biological effects by inducing apoptosis.  相似文献   

16.
Isolated spore coats of a marine Bacillus species were incubated in 25 mM MnCl(2) at pH 7.5. Manganese precipitates, formed on the coat surfaces, were analyzed by transmission electron microscopy, electron diffraction, and energy-dispersive X-ray spectroscopy. Initially, an amorphous manganese oxide was observed on the coats which recrystallized to hausmannite after prolonged incubation in the MnCl(2) solution. The spore coats catalyze the oxidation of Mn(II) and have no structural influence on the final mineral phase precipitated.  相似文献   

17.
Analysis of rodent brains with X‐ray fluorescence (XRF) microscopy combined with immunohistochemistry allowed us to demonstrate that local Cu concentrations are thousands of times higher in the glia of the subventricular zone (SVZ) than in other cells. Using XRF microscopy with subcellular resolution and intracellular X‐ray absorption spectroscopy we determined the copper (I) oxidation state and the sulfur ligand environment. Cu K‐edge X‐ray absorption near edge spectroscopy is consistent with Cu being bound as a multimetallic Cu‐S cluster similar to one present in Cu‐metallothionein. Analysis of age‐related changes show that Cu content in astrocytes of the SVZ increases fourfold from 3 weeks to 9 months, while Cu concentration in other brain areas remain essentially constant. This increase in Cu correlates with a decrease in adult neurogenesis assessed using the Ki67 marker (both, however, can be age‐related effects). We demonstrate that the Cu distribution and age‐related concentration changes in the brain are highly cell specific.  相似文献   

18.
The potential use of the immobilized Mentha arvensis distillation waste (IMADW) biomass for removal and recovery of Cu(II) and Zn(II) from aqueous was evaluated in the present study. Biosorption capacity of Cu(II) and Zn(II) on IMADW increased with increase in pH reaching a maximum at 5 for Cu(II) and 6 for Zn(II). The equilibrium sorption data agreed well with Langmuir isotherm model and pseudo-second-order kinetic model in batch mode. Cu(II) and Zn(II) uptake by IMADW was best described by pseudo-first-order kinetic model in continuous mode. Maximum Cu(II) and Zn(II) uptake by IMADW was 104.48 and 107.75 mg/g, respectively. Fourier Transform Infrared spectroscopy (FTIR) and scanning electron microscopy (SEM) were also carried out to investigate functional groups and surface changes of biomass. The results showed that IMADW biomass is a potential biomaterial to remove Cu(II) and Zn(II) ions with a high biosorption capacity from aqueous solutions.  相似文献   

19.
The effect of Zn(2+) or Cu(2+) ions on Mn-depleted photosystem II (PS II) has been investigated using EPR spectroscopy. In Zn(2+)-treated and Cu(2+)-treated PS II, chemical reduction with sodium dithionite gives rise to a signal attributed to the plastosemiquinone, Q(A)(*)(-), the usual interaction with the non-heme iron being lost. The signal was identified by Q-band EPR spectroscopy which partially resolves the typical g-anisotropy of the semiquinone anion radical. Illumination at 200 K of the unreduced samples gives rise to a single organic free radical in Cu(2+)-treated PS II, and this is assigned to a monomeric chlorophyll cation radical, Chl a(*)(+), based on its (1)H-ENDOR spectrum. The Zn(2+)-treated PS II under the same conditions gives rise to two radical signals present in equal amounts and attributed to the Chl a(*)(+) and the Q(A)(*)(-) formed by light-induced charge separation. When the Cu(2+)-treated PS II is reduced by sodium ascorbate, at >/=77 K electron donation eliminates the donor-side radical leaving the Q(A)(*)(-) EPR signal. The data are explained as follows: (1) Cu(2+) and Zn(2+) have similar effects on PS II (although higher concentrations of Zn(2+) are required) causing the displacement of the non-heme Fe(2+). (2) In both cases chlorophyll is the electron donor at 200 K. It is proposed that the lack of a light-induced Q(A)(*)(-) signal in the unreduced Cu(2+)-treated sample is due to Cu(2+) acting as an electron acceptor from Q(A)(*)(-) at low temperature, forming the Cu(+) state and leaving the electron donor radical Chl a(*)(+) detectable by EPR. (3) The Cu(2+) in PS II is chemically reducible by ascorbate prior to illumination, and the metal can therefore no longer act as an electron acceptor; thus Q(A)(*)(-) is generated by illumination in such samples. (4) With dithionite, both the Cu(2+) and the quinone are reduced resulting in the presence of Q(A)(*)(-) in the dark. The suggested high redox potential of Cu(2+) when in the Fe(2+) site in PS II is in contrast to the situation in the bacterial reaction center where it has been shown in earlier work that the Cu(2+) is unreduced by dithionite. It cannot be ruled out however that Q(A)-Cu(2+) is formed and a magnetic interaction is responsible for the lack of the Q(A)(-) signal when no exogenous reductant is present. With this alternative possibility, the effects of reductants would be explained as the loss of Cu(2+) (due to formation of Cu(+)) leading to loss of the Cu(2+) from the Fe(2+) site due to the binding equilibrium. The quite different binding and redox behavior of the metal in the iron site in PS II compared to that of the bacterial reaction center is presumably a further reflection of the differences in the coordination of the iron in the two systems.  相似文献   

20.
The conditional stability constant at pH 7.4 for Cu(II) binding at the N-terminal site (NTS) of human serum albumin (HSA) was determined directly by competitive UV–vis spectroscopy titrations using nitrilotriacetic acid (NTA) as the competitor in 100 mM NaCl and 100 mM N-(2-hydroxyethyl)piperazine-N′-ethanesulfonic acid (Hepes). The log K NTSc value of 12.0 ± 0.1 was determined for HSA dissolved in 100 mM NaCl. A false log log K NTSc value of 11.4 ± 0.1 was obtained in the 100 mM Hepes buffer, owing to the formation of a ternary Cu(NTA)(Hepes) complex. The impact of the picomolar affinity of HSA for Cu(II) on the availability of these ions in neurodegenerative disorders is briefly discussed.  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