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 311 毫秒
1.
Beta-amyloid peptides (Abetas) bind to several G-protein coupled receptor proteins and stimulate GTPase activity in neurons. In this study we determined the effects of Abeta(1-42), Abeta(1-40), Abeta(25-35) and their mixtures on [(35)S]GTP binding in rat brain cortical membranes in the absence and presence of zinc. We found that the Abetas alone induced a concentration-dependent activation of G-proteins (IC50 approximately 10(-6) m), while aggregated Abeta fibrils only affected GTP binding at concentrations above 10(-5) m. Mixing Abeta(25-35) with Abeta(1-42) or Abeta(1-40) induced a several-fold increase in GTP-binding. This potentiation followed a bell shaped curve with a maximum at 50 : 50 ratios. No potentiating effect could be seen by mixing Abeta(1-40) and Abeta(1-42) or highly aggregated Abetas. Zinc had no effect on Abeta(1-40/42) but strongly potentiated the Abeta(25-35) or the mixed peptides-induced GTP-binding. Changes in secondary structure accompanied the mixed peptides or the peptide/zinc complexes induced potentiation, revealing that structural alterations are behind the increased biological action. These concentration dependent potentiating effects of zinc and the peptide mixtures could be physiologically important at brain regions where peptide fragments and/or zinc are present at elevated concentrations.  相似文献   

2.
Metallothionein-III (MT-III) protects cerebral cortical neurons in established culture from the toxic effect of amyloid beta peptides (Abetas). Protection is concentration dependent and approaches 100% at 0.1 microM. The EC(50) value estimated at 5 microM Abeta(1-40) is 2 nM. At higher concentrations (>0.1 microM), MT-III also antagonizes the trophic effect of Abeta(1-40) on cerebral cortical neurons in early cultures. Because only the fibrillar, SDS-resistant form of Abeta aggregates are thought to be neurotoxic, we analyzed and compared Abeta(1-40) aggregates formed in the presence and absence of MT-III using SDS-PAGE. Results show that aggregates formed in the absence of MT-III are predominantly SDS-resistant whereas those formed in its presence are mostly SDS-soluble. Neither MT-I nor -II exhibits any of the effects of MT-III. On the basis of these results, we propose that MT-III alleviates Abetas' neurotoxic effect by abolishing the formation of toxic aggregates of Abetas and that it may play a specific and important role in protecting the brain from the deleterious effects of Abetas.  相似文献   

3.
Gamma-secretase cleaves the transmembrane domain of beta-amyloid precursor protein at multiple sites referred to as gamma-, epsilon-, and zeta-cleavage sites. We previously showed that N-[N-(3,5-difluorophenacetyl)-L-alanyl]-S-phenylglycine t-butyl ester (DAPT), a potent dipeptide gamma-secretase inhibitor, causes differential accumulation of longer amyloid beta-proteins (Abetas) within Chinese hamster ovary cells co-expressing beta C-terminal fragment and wild-type presenilin 1 (C99/wtPS1 cells). In this study, we used sucrose density gradient centrifugation to fractionate the membranes from C99/wtPS1 cells that had been pretreated with DAPT. We found that accumulating Abeta46 localized exclusively to low density membrane (LDM) domains. Incubating the Abeta46-accumulating LDM domains at 37 degrees C produced Abeta40, Abeta42, Abeta43, and beta-amyloid precursor protein intracellular domain. The addition of L685,458 completely prevented beta-amyloid precursor protein intracellular domain generation and resulted in a large decrease in the level of Abeta46 and the concomitant appearance of Abeta40 and Abeta43 but not Abeta42. Further addition of DAPT suppressed the production of Abeta40/43 and abolished the decrease in the amount of Abeta46. These data indicate that preaccumulated Abeta46 is processed by gamma-secretase to Abeta40/43 but not to Abeta42 in the LDM domains. The amount of newly produced Abeta40 and Abeta43 was roughly equivalent to the decrease in the amount of Abeta46. Temporal profiles did not show a maximal concentration for Abeta43, suggesting that Abeta46 is processed to Abeta40 and Abeta43 through a nonsuccessive process.  相似文献   

4.
Gamma-secretase cleaves the transmembrane domain of beta-amyloid precursor protein at multiple sites. These are referred to as gamma-, zeta-, and epsilon-cleavages. We showed previously that DAPT, a potent dipeptide gamma-secretase inhibitor, caused differential accumulations of longer amyloid beta-proteins (Abetas) (Abeta43 and Abeta46) in CHO cells that are induced to express the beta C-terminal fragment (CTF). To learn more about the cleavage mechanism by gamma-secretase, CHO cell lines coexpressing betaCTF and wild-type or mutant presenilin (PS) 1/2 were generated and treated with DAPT. In all cell lines treated with DAPT, as the levels of Abeta40 decreased, Abeta46 accumulated to varying extents. In wild-type PS1 or M146L mutant PS1 cells, substantial amounts of Abeta43 and Abeta46 accumulated. In contrast, this was not the case with wild-type PS2 cells. In M233T mutant PS1 cells, significant amounts of Abeta46 and Abeta48 accumulated differentially, whereas in N141I mutant PS2 cells, large amounts of Abeta45 accumulated concomitantly with a large decrease in Abeta42 levels. Most interestingly, in G384A mutant PS1 cells, there were no significant accumulations of longer Abetas except for Abeta46. Abeta40 was very susceptible to DAPT, but other Abetas were variably resistant. Complicated suppression and accumulation patterns by DAPT may be explained by stepwise processing of betaCTF from a zeta- or epsilon-cleavage site to a gamma-cleavage site and its preferential suppression of gamma-cleavage over zeta- or epsilon-cleavage.  相似文献   

5.
beta-Amyloid peptides (Abetas) share with lipopolysaccharide, a potent pro-inflammatory agent, the property of stimulating glial cells or macrophages to induce various inflammatory mediators. We recently reported that central administration of lipopolysaccharide induces peripheral interleukin-6 responses via both the central and peripheral norepinephrine system. In this study, the effect of intracerebroventricular injection of various synthetic Abetas on plasma interleukin-6 levels was examined in mice. Abeta(1-42) dose-dependently increased plasma interleukin-6 levels: 'aged' Abeta(1-42) was more effective than fresh, whereas Abeta(42-1) had no effect. 'Aged' Abeta(1-42) (205 pmol/mouse i.c.v.)-induced plasma interleukin-6 peaked at 2 h post injection, which is earlier than the peak time of the Abeta(1-42)-induced brain interleukin-6, tumor necrosis factor-alpha and interleukin-1beta levels, which was 4, 4 and 24 h, respectively. Among various peripheral organs, Abeta(1-42) (205 pmol/mouse i.c.v.) significantly increased interleukin-6 mRNA expression in lymph nodes and liver. Abeta(1-42) (205 pmol/mouse i.c.v.) significantly increased norepinephrine turnover in both hypothalamus and spleen. Either central or peripheral norepinephrine depletion effectively inhibited the Abeta(1-42)-induced peripheral interleukin-6 response. Pretreatment with prazosin (alpha(1)-adrenergic antagonist), yohimbine (alpha(2)-adrenergic antagonist), and ICI-118,551 (beta(2)-adrenergic antagonist), but not with betaxolol (beta(1)-adrenergic antagonist), inhibited Abeta(1-42)-induced plasma interleukin-6 levels. These results demonstrate that centrally administered Abeta(1-42) effectively induces the systemic interleukin-6 response which is mediated, in part, by central Abeta(1-42)-induced activation of the central and the peripheral norepinephrine systems.  相似文献   

6.
Presenilin (PS)/gamma-secretase-mediated intramembranous proteolysis of amyloid precursor protein produces amyloid beta (Abeta) peptides in which Abeta species of different lengths are generated through multiple cleavages at the gamma-, zeta-, and epsilon-sites. An increased Abeta42/Abeta40 ratio is a common characteristic of most cases of familial Alzheimer disease (FAD)-linked PS mutations. However, the molecular mechanisms underlying amyloid precursor protein proteolysis leading to increased Abeta42/Abeta40 ratios still remain unclear. Here, we report our findings on the enzymatic analysis of gamma-secretase derived from I213T mutant PS1-expressing PS1/PS2-deficient (PS(-/-)) cells and from the brains of I213T mutant PS1 knock-in mice. Kinetics analyses revealed that the FAD mutation reduced de novo Abeta generation, suggesting that mutation impairs the total catalytic rate of gamma-secretase. Analysis of each Abeta species revealed that the FAD mutation specifically reduced Abeta40 levels more drastically than Abeta42 levels, leading to an increased Abeta42/Abeta40 ratio. By contrast, the FAD mutation increased the generation of longer Abeta species such as Abeta43, Abeta45, and >Abeta46. These results were confirmed by analyses of gamma-secretase derived from I213T knock-in mouse brains, in which the reduction of de novo Abeta generation was mutant allele dose-dependent. Our findings clearly indicate that the mechanism underlying the increased Abeta42/Abeta40 ratio observed in cases of FAD mutations is related to the differential inhibition of gamma-site cleavage reactions, in which the reaction producing Abeta40 is subject to more inhibition than that producing Abeta42. Our results also provide novel insight into how enhancing the generation of longer Abetas may contribute to Alzheimer disease onset.  相似文献   

7.
The progressive cerebral deposition of a 40-42 residues amyloid beta-peptide (Abeta) is regarded as a major factor in the onset of the Alzheimer's disease. It has recently been shown that Abeta(1-40) is cleaved by Escherichia coli pitrilysin, a homologue of insulysin, at a specific site. To facilitate the studies on a recognition mechanism of Abeta by pitrilysin, an overproduction system of Abeta(1-40) as a fusion protein with E. coli RNase HI was constructed. This fusion protein was designed such that an Abeta(1-40) derivative, Abeta(1-40)*, in which Lys16 and Lys28 of Abeta(1-40) are simultaneously replaced by Ala, is attached to the C-terminus of E. coli RNase HI and Abeta(1-40)* is separated from RNase HI upon cleavage with lysyl endopeptidase. The fusion protein was overproduced in E. coli in inclusion bodies, solubilized and purified in the presence of guanidine hydrochloride, and cleaved by lysyl endopeptidase. Abeta(1-40)* was purified from the resultant peptide fragments by reverse-phase HPLC. Measurement of the far-UV CD spectra suggests that Abeta(1-40)* is conformationally similar to Abeta(1-40). However, the thioflavin T binding assay suggests that Abeta(1-40)* is more amyloidogenic than Abeta(1-40). Nevertheless, Abeta(1-40)* was cleaved by pitrilysin at the site identical to that in Abeta(1-40).  相似文献   

8.
Extracellular amyloid beta peptides (Abetas) have long been thought to be a primary cause of Alzheimer's disease (AD). Now, detection of intracellular neuronal Abeta1--42 accumulation before extracellular Abeta deposits questions the relevance of intracellular peptides in AD. In the present study, we directly address whether intracellular Abeta is toxic to human neurons. Microinjections of Abeta1--42 peptide or a cDNA-expressing cytosolic Abeta1--42 rapidly induces cell death of primary human neurons. In contrast, Abeta1--40, Abeta40--1, or Abeta42--1 peptides, and cDNAs expressing cytosolic Abeta1--40 or secreted Abeta1--42 and Abeta1--40, are not toxic. As little as a 1-pM concentration or 1500 molecules/cell of Abeta1--42 peptides is neurotoxic. The nonfibrillized and fibrillized Abeta1--42 peptides are equally toxic. In contrast, Abeta1--42 peptides are not toxic to human primary astrocytes, neuronal, and nonneuronal cell lines. Inhibition of de novo protein synthesis protects against Abeta1--42 toxicity, indicating that programmed cell death is involved. Bcl-2, Bax-neutralizing antibodies, cDNA expression of a p53R273H dominant negative mutant, and caspase inhibitors prevent Abeta1--42-mediated human neuronal cell death. Taken together, our data directly demonstrate that intracellular Abeta1--42 is selectively cytotoxic to human neurons through the p53--Bax cell death pathway.  相似文献   

9.
In order to develop transgenic animal models that selectively overexpress various Abeta peptides, we have developed a novel expression system that selectively expresses Abeta40 or Abeta42 in the secretory pathway. This system utilizes fusion constructs in which the sequence encoding the 23-amino-acid ABri peptide at the carboxyl terminus of the 266-amino-acid type 2 transmembrane protein BRI is replaced with a sequence encoding either Abeta40 or Abeta42. Constitutive processing of the resultant BRI-Abeta fusion proteins in transfected cells results in high-level expression and secretion of the encoded Abeta peptide. Significantly, expression of Abeta42 from the BRI-Abeta42 construct resulted in no increase in secreted Abeta40, suggesting that the majority of Abeta42 is not trimmed by carboxypeptidase to Abeta40 in the secretory pathway.  相似文献   

10.
Antibodies against APP, a precursor of Abeta deposited in Alzheimer's disease brain, have been shown to cause neuronal death. Therefore, it is important to determine whether Abeta mediates antibody-induced neurotoxicity. When primary neurons were treated with anti-APP antibodies, Abeta40 and Abeta42 in the cultured media were undetectable by an assay capable of detecting 100 nM Abeta peptides. However, exogenously treated Abeta1-42 or Abeta1-43 required >3 microM to exert neurotoxicity, and 25 microM Abeta1-40 was not neurotoxic. Glutathione-ethyl-ester inhibited neuronal death by anti-APP antibody, but not death by Abeta1-42, whereas serum attenuated toxicity by Abeta1-42, but not by anti-APP antibody. Using immortalized neuronal cells, we specified the domain responsible for toxicity to be cytoplasmic His(657)-Lys(676), but not the Abeta1-42 region, of APP. This indicates that neuronal cell death by anti-APP antibody is not mediated by secreted Abeta.  相似文献   

11.
We previously found that pathophysiological concentrations (< or = 10 nm) of an amyloid beta protein (Abeta25-35) reduced the plasma membrane phosphatidylinositol monophosphate level in cultured rat hippocampal neurons with a decrease in phosphatidylinositol 4-monophosphate-dependent Cl- -ATPase activity. As this suggested an inhibitory effect of Abeta25-35 on plasma membrane phosphatidylinositol 4-kinase (PI4K) activity, in vitro effects of Abetas on PI4K activity was examined using rat brain subcellular fractions and recombinant human type II PI4K (PI4KII). Abeta25-35 (10 nm) inhibited PI4KII activity, but neither PI 3-kinase (PI3K) nor type III PI4K (PI4KIII) activity, in microsomal fractions, while 100 nm Abeta25-35 inhibited PI3K activity in mitochondrial fractions. In plasma membrane-rich fractions, Abetas (> 0.5 nm) dose-dependently inhibited PI4KII activity, the maximal inhibition to 77-87% of control being reached around 10 nm of Abetas without significant changes in apparent Km values for ATP and PI, suggesting non-competitive inhibition by Abetas. The inhibition by 10 nm Abeta25-35 was reversible. In recombinant human PI4KIIalpha, inhibition profiles of Abetas were similar to those in rat brain plasma membranes. Therefore, pathophysiological concentrations of Abetas directly and reversibly inhibited plasma membrane PI4KII activity, suggesting that plasma membrane PI4KII is a target of Abetas in the pathogenesis of Alzheimer's disease.  相似文献   

12.
We investigated changes in cytosolic phospholipase A(2) (cPLA(2)) and calcium-independent PLA(2) (iPLA(2)) activities in bovine retina capillary pericytes after stimulation with 50 microM amyloid-beta (Abeta) (1-42) and its (25-35) fragment, over 24 h (mild, sublethal model of cell damage). In the presence of Abeta peptides, we found that cPLA(2) activity was increased and translocated from the cytosolic fraction to the membrane system, particularly in the nuclear region. Reversed-sequence Abeta(35-25) peptide did not stimulate or induce cPLA(2) translocation. Exposure to both Abeta peptides had no significant effect on cPLA(2) protein content as tested by Western immunoblot analysis. The addition of Abetas to quiescent pericytes was followed by phosphorylation of cPLA(2) and arachidonic acid release. Treatment with inhibitors (AACOCF(3), staurosporine and cycloheximide) resulted in a sharp decrease in basal and stimulated cPLA(2) activity. Inactivating effects of bromoenol lactone (BEL), inhibitor of iPLA(2), demonstrated that the stimulation of total PLA(2) activity by Abetas was mediated by both PLA(2) enzymes. Taken together with our previous observations that both Abeta peptides may induce hydrolysis of phosphatidylcholine, the present results provide evidence that this process is cooperatively mediated by cPLA(2) activation/translocation and iPLA(2) activation. The effect is very likely triggered by a mild prooxidant mechanism which was not able to divert the cell to degeneration. The data confirm the hypothesis that pericytes could be a target of potential vascular damage and reactivity during processes involving amyloid accumulation.  相似文献   

13.
Zhang L  Song L  Terracina G  Liu Y  Pramanik B  Parker E 《Biochemistry》2001,40(16):5049-5055
Recent studies of gamma-secretase have pointed out that it may be comprised of a multisubunit complex with presenilin 1 and presenilin 2 as central components. Elucidation of the biochemical mechanism of this enzymatic activity will provide important information for developing gamma-secretase inhibitors in Alzheimer's disease therapy. Here we describe the biochemical characterization of gamma-secretase activities using a sensitive, membrane-based assay system. Membranes were isolated from 293 cells expressing C99, the substrate of gamma-secretase. Upon incubation at 37 degrees C, C99 is cleaved by the endogenous gamma-secretase, and Abeta peptides are liberated. Abeta40 and Abeta42 gamma-secretase activities are very similar in terms of their kinetic profiles and pH dependence, supporting the notion that a single enzyme is involved in both Abeta40 and Abeta42 production. Pepstatin A inhibited Abeta40 and Abeta42 gamma-secretase activities with similar potency. Peptide difluoroketone and peptide aldehyde inhibitors inhibited Abeta40 production in a dose-dependent fashion, enhanced Abeta42 production at low concentrations, and inhibited Abeta42 production at high concentrations. Although the selective increase of Abeta42 by low concentrations of peptide difluoroketone and peptide aldehyde inhibitors has been reported in intact cells, the finding that this phenomenon occurs in a membrane-based assay system suggests that these compounds increase Abeta42 by a direct effect on gamma-secretase. The ability of these compounds to increase Abeta42 production may reflect allosteric modulation of the gamma-secretase complex by a mechanism related to that responsible for the increase of Abeta42 production by mutations in presenilins.  相似文献   

14.
Strong support for a primary causative role of the Abeta peptides in the development of Alzheimer's disease (AD) neurodegeneration derives from reports that presenilin familial AD (FAD) mutants alter amyloid precursor protein processing, thus increasing production of neurotoxic Abeta 1-42 (Abeta 42). This effect of FAD mutants is also reflected in an increased ratio of peptides Abeta 42 over Abeta 1-40 (Abeta 40). In the present study, we show that several presenilin 1 FAD mutants failed to increase production of Abeta 42 or the Abeta 42/40 ratio. Our data suggest that the mechanism by which FAD mutations promote neurodegeneration and AD may be independent of their effects on Abeta production.  相似文献   

15.
We investigated the effects of different apolipoprotein E (apoE) isoforms, Abeta (1-42), and apoE/Abeta complexes on PKC-alpha translocation and APP processing in human SH-SY5Y neuroblastoma cells and fibroblasts. Treatment of cells with either 10 nM apoE3 or apoE4, 10 microM Abeta (1-42), or apoE/Abeta complexes induced significant translocation of PKC-alpha in both cell types. Effects were seen using both human recombinant apoE and apoE loaded into beta-very low density lipoprotein (beta-VLDL) particles. Time course (5-24 h) studies of APP processing revealed that some conditions induced transient or moderate increases in the secretion of proteins detected by 22C11. In contrast, the secretion of alpha-secretase cleaved APP was either not modified or transiently decreased, as determined by immunoblotting with the antibody 6E10. These results suggest that apoE, Abeta (1-42) and apoE/Abeta complexes can modulate PKC activity but do not have major consequences for APP processing. These effects could contribute to the reported PKC alterations seen in AD. However, it is unlikely that the contribution of different apoE isoforms to AD pathology occurs via effects on APP processing.  相似文献   

16.
beta-Amyloid peptides (Abeta) that form the senile plaques of Alzheimer disease consist mainly of 40- and 42-amino acid (Abeta 40 and Abeta 42) peptides generated from the cleavage of the amyloid precursor protein (APP). Generation of Abeta involves beta-secretase and gamma-secretase activities and is regulated by membrane trafficking of the proteins involved in Abeta production. Here we describe a new small molecule, EHT 1864, which blocks the Rac1 signaling pathways. In vitro, EHT 1864 blocks Abeta 40 and Abeta 42 production but does not impact sAPPalpha levels and does not inhibit beta-secretase. Rather, EHT 1864 modulates APP processing at the level of gamma-secretase to prevent Abeta 40 and Abeta 42 generation. This effect does not result from a direct inhibition of the gamma-secretase activity and is specific for APP cleavage, since EHT 1864 does not affect Notch cleavage. In vivo, EHT 1864 significantly reduces Abeta 40 and Abeta 42 levels in guinea pig brains at a threshold that is compatible with delaying plaque accumulation and/or clearing the existing plaque in brain. EHT 1864 is the first derivative of a new chemical series that consists of candidates for inhibiting Abeta formation in the brain of AD patients. Our findings represent the first pharmacological validation of Rac1 signaling as a target for developing novel therapies for Alzheimer disease.  相似文献   

17.
The concentration of beta-amyloid peptide (Abeta), x-42 or x-40 amino acids long, increases in brain with the progression Alzheimer's disease (AD). These peptides are deposited extracellularly as highly insoluble fibrils that form densities of amyloid plaques. Abeta fibrillization is a complex polymerization process preceded by the formation of oligomeric and prefibrillar Abeta intermediates. In some of our in vitro studies, in which the kinetics of intermediate steps of fibril formation were examined, we used concentrations of synthetic Abeta that exceed what is normally employed in fibrillization studies, 300-600 microM. At these concentrations, in a cell free system and under physiological conditions, Abeta 1-40 peptide (Abeta40) forms fibrils that spontaneously assemble into clearly defined spheres, "betaamy balls", with diameters of approximately 20-200 microm. These supramolecular structures show weak birefringence with Congo red staining and high stability with prolonged incubation times (at least 2 weeks) at 30 degrees C, freezing, and dilution in H(2)O. At 600 microM, they are detected after incubation for approximately 20 h. Abeta peptide 1-42 (Abeta42) lacks the ability to form betaamy balls but accelerates Abeta40 betaamy ball formation at low stoichiometric levels (1:20 Abeta42:Abeta40 ratio). Abeta42 levels above this (=10-50% w/w) impede Abeta40 betaamy ball formation. Using light (LM) and electron microscopy (EM), this study examines the gross morphology and ultrastructure of Abeta40 betaamy balls and their time course of formation, in the absence and presence of Abeta42, along with some stability measures. As spheres of a misfolded protein, betaamy balls resemble both AD Abeta senile plaques and neuronal inclusion bodies associated with other neurodegenerative diseases.  相似文献   

18.
To investigate the significance of endoproteolytic processing of presenilin 2 (PS2) on its pathological function, we constructed PS2 cDNAs causing amino acid substitutions or deletions around the cleavage site. We found that a PS2 mutant (Del3) with a 20-amino acid deletion was not endoproteolytically processed, while other PS2s with amino acid substitutions and short deletions were cleaved. Overproduction of all the mutant proteins led to a compensatory decrease of endogenous PS1 fragments, but did not affect the amyloid beta peptide X-42/Abeta X-40 ratio without the familial Alzheimer's disease mutation. The Del3 mutant did not exhibit significant deficits in gamma-secretase activity. The turnover rate of the Del3 holoprotein was the same as that of full-length PS2. These data suggest that the determinants of the PS2 cleavage site reside within a large region and that the pathological function of PS2 is exerted by familial Alzheimer's disease mutations not related to the cleavage of holoproteins. We also found that PS2 with an 18-amino acid deletion at the C-terminal end was not processed. Its overexpression led neither to diminished accumulation of endogenous PS1 fragments nor to increased production of amyloid beta peptide X-42. The C-terminal end of PS2 seems to possess the signal for entry into the processing pathway.  相似文献   

19.
Zinc, iron and copper are concentrated in senile plaques of Alzheimer disease. Copper and iron catalyze the Fenton-Haber-Weiss reaction, which likely contributes to oxidative stress in neuronal cells. In this study, we found that ascorbate oxidase activity and the intensity of ascorbate radicals measured using ESR spectroscopy, generated by free Cu(II), was decreased in the presence of amyloid-beta (Abeta), the major component of senile plaques. Specifically, the ascorbate oxidase activity was strongly inhibited (85% decrease) in the presence of Abeta1-16 or Abeta1-42, whereas it was only slightly inhibited in the presence of Abeta1-12 or Abeta25-35 (<20% inhibition). Ascorbate-dependent hydroxyl radical generation by free Cu(II) decreased in the presence of Abeta in the identical order of Abeta1-42, Abeta1-16 > Abeta1-12 and was abolished in the presence of 2-fold molar excess glycylhystidyllysine (GHK). Ascorbate oxidase activity and ascorbate-dependent hydroxyl radical generation by free Fe(III) were inhibited by Abeta1-42, Abeta1-16, and Abeta1-12. Although Cu(II)-Abeta shows a significant SOD-like activity, the rate constant for the reaction of superoxide with Cu(II)-Abeta was much slower than that with SOD. Overall, our results suggest that His6, His13, and His14 residues of Abeta1-42 control the redox activity of transition metals present in senile plaques.  相似文献   

20.
Urea-based beta-amyloid (Abeta) SDS-polyacrylamide gel electrophoresis and immunoblots were used to analyze the generation of Abeta peptides in conditioned medium from primary mouse neurons and a neuroglioma cell line, as well as in human cerebrospinal fluid. A comparable and highly conserved pattern of Abeta peptides, namely, 1-40/42 and carboxyl-terminal-truncated 1-37, 1-38, and 1-39, was found. Besides Abeta1-42, we also observed a consistent elevation of amino-terminal-truncated Abeta2-42 in a detergent-soluble pool in brains of subjects with Alzheimer's disease. Abeta2-42 was also specifically elevated in cerebrospinal fluid samples of Alzheimer's disease patients. To decipher the contribution of potential different gamma-secretases (presenilins (PSs)) in generating the amino-terminal- and carboxyl-terminal-truncated Abeta peptides, we overexpressed beta-amyloid precursor protein (APP)-trafficking mutants in PS1+/+ and PS1-/- neurons. As compared with APP-WT (primary neurons from control or PS1-deficient mice infected with Semliki Forest virus), PS1-/- neurons and PS1+/+ neurons overexpressing APP-Deltact (a slow-internalizing mutant) show a decrease of all secreted Abeta peptide species, as expected, because this mutant is processed mainly by alpha-secretase. This drop is even more pronounced for the APP-KK construct (APP mutant carrying an endoplasmic reticulum retention motif). Surprisingly, Abeta2-42 is significantly less affected in PS1-/- neurons and in neurons transfected with the endocytosis-deficient APP-Deltact construct. Our data confirm that PS1 is closely involved in the production of Abeta1-40/42 and the carboxyl-terminal-truncated Abeta1-37, Abeta1-38, and Abeta1-39, but the amino-terminal-truncated and carboxyl-terminal-elongated Abeta2-42 seems to be less affected by PS1 deficiency. Moreover, our results indicate that the latter Abeta peptide species could be generated by a beta(Asp/Ala)-secretase activity.  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