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.
Miura T  Suzuki K  Kohata N  Takeuchi H 《Biochemistry》2000,39(23):7024-7031
Aggregation of the amyloid beta-peptide (Abeta) into insoluble fibrils is a key pathological event in Alzheimer's disease. Zn(II) induces the Abeta aggregation at acidic-to-neutral pH, while Cu(II) is an effective inducer only at mildly acidic pH. We have examined Zn(II) and Cu(II) binding modes of Abeta and their pH dependence by Raman spectroscopy. The Raman spectra clearly demonstrate that three histidine residues in the N-terminal hydrophilic region provide primary metal binding sites and the solubility of the metal-Abeta complex is correlated with the metal binding mode. Zn(II) binds to the N(tau) atom of the histidine imidazole ring and the peptide aggregates through intermolecular His(N(tau))-Zn(II)-His(N(tau)) bridges. The N(tau)-metal ligation also occurs in Cu(II)-induced Abeta aggregation at mildly acidic pH. At neutral pH, however, Cu(II) binds to N(pi), the other nitrogen of the histidine imidazole ring, and to deprotonated amide nitrogens of the peptide main chain. The chelation of Cu(II) by histidine and main-chain amide groups results in soluble Cu(II)-Abeta complexes. Under normal physiological conditions, Cu(II) is expected to protect Abeta against Zn(II)-induced aggregation by competing with Zn(II) for histidine residues of Abeta.  相似文献   

2.
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   

3.
There is evidence that binding of metal ions like Zn2+ and Cu2+ to amyloid beta-peptides (Abeta) may contribute to the pathogenesis of Alzheimer's disease. Cu2+ and Zn2+ form complexes with Abeta peptides in vitro; however, the published metal-binding affinities of Abeta vary in an enormously large range. We studied the interactions of Cu2+ and Zn2+ with monomeric Abeta(40) under different conditions using intrinsic Abeta fluorescence and metal-selective fluorescent dyes. We showed that Cu(2+) forms a stable and soluble 1 : 1 complex with Abeta(40), however, buffer compounds act as competitive copper-binding ligands and affect the apparent K(D). Buffer-independent conditional K(D) for Cu(II)-Abeta(40) complex at pH 7.4 is equal to 0.035 micromol/L. Interaction of Abeta(40) with Zn2+ is more complicated as partial aggregation of the peptide occurs during zinc titration experiment and in the same time period (within 30 min) the initial Zn-Abeta(40) complex (K(D) = 60 micromol/L) undergoes a transition to a more tight complex with K(D) approximately 2 micromol/L. Competition of Abeta(40) with ion-selective fluorescent dyes Phen Green and Zincon showed that the K(D) values determined from intrinsic fluorescence of Abeta correspond to the binding of the first Cu2+ and Zn2+ ions to the peptide with the highest affinity. Interaction of both Zn2+ and Cu2+ ions with Abeta peptides may occur in brain areas affected by Alzheimer's disease and Zn2+-induced transition in the peptide structure might contribute to amyloid plaque formation.  相似文献   

4.
The abnormal form of the prion protein (PrP) is believed to be responsible for the transmissible spongiform encephalopathies. A peptide encompassing residues 106-126 of human PrP (PrP106-126) is neurotoxic in vitro due its adoption of an amyloidogenic fibril structure. The Alzheimer's disease amyloid beta peptide (Abeta) also undergoes fibrillogenesis to become neurotoxic. Abeta aggregation and toxicity is highly sensitive to copper, zinc, or iron ions. We show that PrP106-126 aggregation, as assessed by turbidometry, is abolished in Chelex-100-treated buffer. ICP-MS analysis showed that the Chelex-100 treatment had reduced Cu(2+) and Zn(2+) levels approximately 3-fold. Restoring Cu(2+) and Zn(2+) to their original levels restored aggregation. Circular dichroism showed that the Chelex-100 treatment reduced the aggregated beta-sheet content of the peptide. Electron paramagnetic resonance spectroscopy identified a 2N1S1O coordination to the Cu(2+) atom, suggesting histidine 111 and methionine 109 or 112 are involved. Nuclear magnetic resonance confirmed Cu(2+) and Zn(2+) binding to His-111 and weaker binding to Met-112. An N-terminally acetylated PrP106-126 peptide did not bind Cu(2+), implicating the free amino group in metal binding. Mutagenesis of either His-111, Met-109, or Met-112 abolished PrP106-126 neurotoxicity and its ability to form fibrils. Therefore, Cu(2+) and/or Zn(2+) binding is critical for PrP106-126 aggregation and neurotoxicity.  相似文献   

5.
T Miura  A Hori-i  H Mototani  H Takeuchi 《Biochemistry》1999,38(35):11560-11569
The cellular form of prion protein is a precursor of the infectious isoform, which causes fatal neurodegenerative diseases through intermolecular association. One of the characteristics of the prion protein is a high affinity for Cu(II) ions. The site of Cu(II) binding is considered to be the N-terminal region, where the octapeptide sequence PHGGGWGQ repeats 4 times in tandem. We have examined the Cu(II) binding mode of the octapeptide motif and its pH dependence by Raman and absorption spectroscopy. At neutral and basic pH, the single octapeptide PHGGGWGQ forms a 1:1 complex with Cu(II) by coordinating via the imidazole N pi atom of histidine together with two deprotonated main-chain amide nitrogens in the triglycine segment. A similar 1:1 complex is formed by each octapeptide unit in (PHGGGWGQ)2 and (PHGGGWGQ)4. Under weakly acidic conditions (pH approximately 6), however, the Cu(II)-amide- linkages are broken and the metal binding site of histidine switches from N pi to N tau to share a Cu(II) ion between two histidine residues of different peptide chains. The drastic change of the Cu(II) binding mode on going from neutral to weakly acidic conditions suggests that the micro-environmental pH in the brain cell regulates the Cu(II) affinity of the prion protein, which is supposed to undergo pH changes in the pathway from the cell surface to endosomes. The intermolecular His(N tau)-Cu(II)-His(N tau) bridge may be related to the aggregation of prion protein in the pathogenic form.  相似文献   

6.
Amyloid beta peptide (Abeta) is the major constituent of extracellular plaques and perivascular amyloid deposits, the pathognomonic neuropathological lesions of Alzheimer's disease. Cu(2+) and Zn(2+) bind Abeta, inducing aggregation and giving rise to reactive oxygen species. These reactions may play a deleterious role in the disease state, because high concentrations of iron, copper, and zinc have been located in amyloid in diseased brains. Here we show that coordination of metal ions to Abeta is the same in both aqueous solution and lipid environments, with His(6), His(13), and His(14) all involved. At Cu(2+)/peptide molar ratios >0.3, Abeta coordinated a second Cu(2+) atom in a highly cooperative manner. This effect was abolished if the histidine residues were methylated at N(epsilon)2, indicating the presence of bridging histidine residues, as found in the active site of superoxide dismutase. Addition of Cu(2+) or Zn(2+) to Abeta in a negatively charged lipid environment caused a conformational change from beta-sheet to alpha-helix, accompanied by peptide oligomerization and membrane penetration. These results suggest that metal binding to Abeta generated an allosterically ordered membrane-penetrating oligomer linked by superoxide dismutase-like bridging histidine residues.  相似文献   

7.
There is now direct evidence that copper is bound to amyloid-beta peptide (Abeta) in senile plaque of Alzheimer's disease. Copper is also linked with the neurotoxicity of Abeta and free radical damage, and Cu(2+) chelators represent a possible therapy for Alzheimer's disease. We have therefore used a range of complementary spectroscopies to characterize the coordination of Cu(2+) to Abeta in solution. The mode of copper binding is highly pH-dependent. EPR spectroscopy indicates that both coppers have axial, Type II coordination geometry, square-planar or square-pyramidal, with nitrogen and oxygen ligands. Circular dichroism studies indicate that copper chelation causes a structural transition of Abeta. Competition studies with glycine and l-histidine indicate that copper binds to Abeta-(1-28) at pH 7.4 with an affinity of K(a) approximately 10(7) m(-1). (1)H NMR indicates that histidine residues are involved in Cu(2+) coordination but that Tyr(10) is not. Studies using analogues of Abeta-(1-28) in which each of the histidine residues have been replaced by alanine or in which the N terminus is acetylated suggest that the N terminus and His(13) are crucial for Cu(2+) binding and that His(6) and His(14) are also implicated. Evidence for the link between Alzheimer's disease and Cu(2+) is growing, and our studies have made a significant contribution to understanding the mode of Cu(2+) binding to Abeta in solution.  相似文献   

8.
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   

9.
Amyloid beta (Abeta) is a central characteristic of Alzheimer's disease (AD). Currently, there is a long-standing dispute regarding the role of Abeta-metal ion (Zn, Cu, and Fe) complexes in AD pathogenesis. Here, we aim to decipher the connection between oxidative damage implicated in AD and Abeta-metal ion complexes. For this purpose we study, using ESR, the modulation of Cu/Fe-induced H 2O 2 decomposition by Abeta 1-28 (Abeta 28), a soluble model of Abeta 40/42. The addition of H 2O 2 to 0.6 nM-360 microM Abeta 28 solutions containing 100 microM Cu(II)/Cu(I)/Fe(II) at pH 6.6 results in a concentration-dependent sigmoidal decay of [*OH] with IC 50 values of 61, 59, and 84 microM, respectively. Furthermore, Abeta 28 reduces 90% of *OH production rate in the Cu(I)-H 2O 2 system in 5 min. Unlike soluble Abeta 28, Abeta 28-Cu aggregates exhibit poor antioxidant activity. The mode of antioxidant activity of soluble Abeta 28 is twofold. The primary (rapid) mechanism involves metal chelation, whereas the secondary (slow) mechanism involves (*)OH scavenging and oxidation of Cu(Fe)-coordinating ligands. On the basis of our findings, we propose that soluble Abeta may play a protective role in the early stages of AD, but not in healthy individuals, where Abeta's concentration is nanomolar. Yet, when Abeta-metal ion complexes undergo aggregation, they significantly lose their protective function and allow oxidative damage to occur.  相似文献   

10.
The binding of Cu(II) to the prion protein is investigated by computations at the B3LYP level of theory on models of the octarepeat domain of the prion protein. The models incorporate the functionality of the glycine (G) and histidine (H) residues which occur in the octarepeat domain, PHGGGWGQ. The copper complexes are designated Cu[HG] and Cu[HGGG]. Coordination to the metal via the imidazole ring of the histidine, the amide carbonyl groups, and the backbone nitrogen atom of the amide groups were examined, as well as several protonation/deprotonation states of each structure. EPR and CD titration experiments suggest that the octarepeat segments of the unstructured N-terminal domain of prion protein can bind Cu(II) in a 1:1 Cu-to-octarepeat ratio. The results identify the extent to which the Cu(II) facilitates peptide backbone deprotonation, and the propensity of binding in the forward (toward the C-terminus) direction from the anchoring histidine residue. A plausible mechanism is suggested for changing from amide O-atom to deprotonated amide N-atom coordination, and for assembly of the observed species in solutions of Cu[PrP] and truncated models of it. A structure is proposed which has the N2O2 coordination pattern for the minor component observed experimentally by EPR spectroscopy for the Cu[HGGG] model. The most stable neutral Cu[HGGG] structure found, with coordination environment N3O1, corresponds to that observed for Cu[HGGGW] and Cu[HGGG] both in the solid state and as the major component in solution at neutral pH.  相似文献   

11.
With a combination of complementary experimental techniques, namely sedimentation assay, Fourier transform infrared spectroscopy, and x-ray absorption spectroscopy, we are able to determine the atomic structure around the metal-binding site in samples where amyloid-beta (Abeta) peptides are complexed with either Cu(II) or Zn(II). Exploiting information obtained on a selected set of fragments of the Abeta peptide, we identify along the sequence the histidine residues coordinated to the metal in the various peptides we have studied (Abeta(1-40), Abeta(1-16), Abeta(1-28), Abeta(5-23), and Abeta(17-40)). Our data can be consistently interpreted assuming that all of the peptides encompassing the minimal 1-16 amino acidic sequence display a copper coordination mode that involves three histidines (His(6), His(13), and His(14)). In zinc-Abeta complexes, despite the fact that the metal coordination appears to be more sensitive to solution condition and shows a less rigid geometry around the binding site, a four-histidine coordination mode is seen to be preferred. Lacking a fourth histidine along the Abeta peptide sequence, this geometrical arrangement hints at a Zn(II)-promoted interpeptide aggregation mode.  相似文献   

12.
Binding of Zn(II), Cu(II) and Fe(II) ions to A beta1-40, A beta1-42 and a single tryptophan mutant of Abeta 1-40 in solution at pH 7.4 was studied by fluorescent titration. Job plots and fitting of titration curves revealed formation of 1:1 and 1:2 peptide-metal complexes. For dimeric peptides A beta1-40 and A betaF4W the order of metal to peptide affinities is Fe < Cu > Zn, which is in agreement with the Irving-Williams series of complex stability. The affinity of A beta1-42 for Fe increases dramatically upon aggregation: K(D) changes from ca. 100 to ca. 0.2 microM.  相似文献   

13.
Transition metals have been frequently recognized as risk factors in neurodegenerative disorders, and brain lesions associated with Alzheimer's disease are rich in Fe(III), Zn(II), and Cu(II). By using different biophysical techniques (nuclear magnetic resonance, circular dichroism, light scattering, and microcalorimetry), we have structurally characterized the binding of Cu(II) to a 198 amino acid fragment of the protein Tau that can mimic both the aggregation behavior and microtubule binding properties of the full-length protein. We demonstrate that Tau can specifically bind one Cu(II) ion per monomer with a dissociation constant in the micromolar range, an affinity comparable to the binding of Cu(II) to other proteins involved in neurodegenerative diseases. NMR spectroscopy showed that two short stretches of residues, (287)VQSKCGS (293) and (310)YKPVDLSKVTSKCGS (324), are primarily involved in copper binding, in agreement with mutational analysis. According to circular dichroism and NMR spectroscopy, Tau remains largely disordered upon binding to Cu(II), although a limited amount of aggregation is induced.  相似文献   

14.
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   

15.
In order to study the metal ion binding of optically active beta-(2-pyridyl)-alpha-alanine, NH2CH(CH2C5H4N)CO2H, Pyala, and beta-(6-methyl-2-pyridyl)-alpha-alanine, NH2CH(CH2C5H3NCH3)CO2H, Mepyala, these pyridine analogs of histidine were synthesized and resolved; absolute configurations were determined for the isolated enatiomers. Protonation constants and formation constants for the binding of L-Pyala, D,L-Pyala, D-Mepyala, and D,L-Mepyala with Cu(II), Ni(II), Co(II), and Zn(II) were determined by potentiometrictitration. They show that the formation constant (Kx) for the reaction, M(L-Pyala) + D-Pyala in equilibrium M(L-Pyala)(D-pyala), is larger (up to 8.7 times larger) than that (K2) for the coordination of the same enantiomer, M(L-Pyala) + L-Pyala in equilibrium M(L-Pyala)2. Although the difference between Kx and K2 can be explained on a statistical basis for Cu(II), the larger differences for Ni(II), Co(II), and Zn(II) demonstrate that their M(L-Pyala)complexes bind D-Pyala more favorably than they do L-Pyala. A similar trend was found for the optical isomers of Mepyala except that the stereoselectivity is less than in the Pyala system. This presumably results from the reduced coordinating ability of the 6-methylpyridyl group as compared to the less crowded pyridyl donor. Using ideas previously applied to the analogous histidine system, the observed stereoselectivities may be explained in terms of the structures of the complexes.  相似文献   

16.
Mounting evidence has shown that dyshomeostasis of the redox-active biometals such as Cuand Fe can lead to oxidative stress,which plays a key role in the neuropathology of Alzheimer's disease(AD).Here we demonstrate that with the formation of Cu(Ⅱ)·Aβ1-40 complexes,copper markedly potentiatesthe neurotoxicity exhibited by β-amyloid peptide (Aβ).A greater amount of hydrogen peroxide was releasedwhen Cu(Ⅱ)·Aβ1-40 complexes was added to the xanthine oxidase/xanthine system detected by potassiumiodide spectrophotometry.Copper bound to Aβ1-40 was observed by electron paramagnetic resonance(EPR) spectroscopy.Circular dichroism (CD) studies indicated that copper chelation could cause a structuraltransition of Aβ.The addition of copper to Aβ introduced an increase on β-sheet as well as α-helix,whichmay be responsible for the aggregation of Aβ.We hypothesized that Aβ aggregation induced by copper maybe responsible for local injury in AD.The interaction between Cu~(2 ) and Aβ also provides a possible mechanismfor the enrichment of metal ions in amyloid plaques in the AD brain.  相似文献   

17.
Aberrant interactions of copper and zinc ions with the amyloid-beta peptide (Abeta) potentiate Alzheimer's disease (AD) by participating in the aggregation process of Abeta and in the generation of reactive oxygen species (ROS). The ROS production and the neurotoxicity of Abeta are associated with copper binding. Metallothionein-3 (Zn(7)MT-3), an intra- and extracellularly occurring metalloprotein, is highly expressed in the brain and downregulated in AD. This protein protects, by an unknown mechanism, cultured neurons from the toxicity of Abeta. Here, we show that a metal swap between Zn(7)MT-3 and soluble and aggregated Abeta(1-40)-Cu(II) abolishes the ROS production and the related cellular toxicity. In this process, copper is reduced by the protein thiolates forming Cu(I)(4)Zn(4)MT-3, in which an air-stable Cu(I)(4)-thiolate cluster and two disulfide bonds are present. The discovered protective effect of Zn(7)MT-3 from the copper-mediated Abeta(1-40) toxicity may lead to new therapeutic strategies for treating AD.  相似文献   

18.
All inherited forms of human prion diseases are linked with mutations in the prion protein (PrP) gene. Here we have investigated the stability and Cu(II) binding properties of three recombinant variants of murine full-length PrP(23-231)-containing destabilizing point mutations that are associated with human Gerstmann-Str?ussler-Scheinker disease (F198S), Creutzfeld-Jakob disease (E200K), and fatal familial insomnia (D178N) by electron paramagnetic resonance and circular dichroism spectroscopy. Furthermore, we analyzed the variants H140S, H177S, and H187S of the isolated C-terminal domain of murine PrP, mPrP(121-231), to test a role of the histidine residues in Cu(II) binding. The F198S and E200K variants of PrP(23-231) differed in Cu(II) binding from the wild-type mPrP(23-231). However, circular dichroism spectroscopy indicated that the variants and the wild type did not undergo conformational changes in the presence of Cu(II). The D178N variant showed a high tendency to aggregate at pH 7.4 both with and without Cu(II). At lower pH values, it showed the same Cu(II) binding behavior as the wild type. The analysis allowed for a better location of the Cu(II) binding sites in the C-terminal part of the protein. Our present data indicate that hereditary forms of prion diseases cannot be rationalized on the basis of altered Cu(II) binding or mutation-induced protein destabilization alone.  相似文献   

19.
The N-terminal metal binding extension of the Cu,Zn superoxide dismutase from Haemophilus ducreyi is constituted by a histidine-rich region followed by a methione-rich sequence which shows high similarity with protein motifs involved in the binding of Cu(I). X-ray absorption spectroscopy experiments selectively carried out with peptides corresponding to the two metal binding regions indicate that both sequences can bind either Cu(II) or Cu(I). However, competition experiments demonstrate that Cu(II) is preferred by histidine residues belonging to the first half of the motif, while the methionine-rich region preferentially binds Cu(I) via the interaction with three methionine sulfur atoms. Moreover, we have observed that the rate of copper transfer from the peptides to the active site of a copper-free form of the Cu,Zn superoxide dismutase mutant lacking the N-terminal extension depends on the copper oxidation state and on the residues involved in metal binding, histidine residues being critically important for the efficient transfer. Differences in the enzyme reactivation rates in the presence of mixtures of the two peptides when compared to those obtained with the single peptides suggest that the two halves of the N-terminal domain functionally interact during the process of copper transfer, possibly through subtle modifications of the copper coordination environment.  相似文献   

20.
Cu(II) ions are implicated in the pathogenesis of Alzheimer disease by influencing the aggregation of the amyloid-β (Aβ) peptide. Elucidating the underlying Cu(II)-induced Aβ aggregation is paramount for understanding the role of Cu(II) in the pathology of Alzheimer disease. The aim of this study was to characterize the qualitative and quantitative influence of Cu(II) on the extracellular aggregation mechanism and aggregate morphology of Aβ(1-40) using spectroscopic, microelectrophoretic, mass spectrometric, and ultrastructural techniques. We found that the Cu(II):Aβ ratio in solution has a major influence on (i) the aggregation kinetics/mechanism of Aβ, because three different kinetic scenarios were observed depending on the Cu(II):Aβ ratio, (ii) the metal:peptide stoichiometry in the aggregates, which increased to 1.4 at supra-equimolar Cu(II):Aβ ratio; and (iii) the morphology of the aggregates, which shifted from fibrillar to non-fibrillar at increasing Cu(II):Aβ ratios. We observed dynamic morphological changes of the aggregates, and that the formation of spherical aggregates appeared to be a common morphological end point independent on the Cu(II) concentration. Experiments with Aβ(1-42) were compatible with the conclusions for Aβ(1-40) even though the low solubility of Aβ(1-42) precluded examination under the same conditions as for the Aβ(1-40). Experiments with Aβ(1-16) and Aβ(1-28) showed that other parts than the Cu(II)-binding His residues were important for Cu(II)-induced Aβ aggregation. Based on this study we propose three mechanistic models for the Cu(II)-induced aggregation of Aβ(1-40) depending on the Cu(II):Aβ ratio, and identify key reaction steps that may be feasible targets for preventing Cu(II)-associated aggregation or toxicity in Alzheimer disease.  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