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 15 毫秒
1.
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   

2.
Lim KH  Kim YK  Chang YT 《Biochemistry》2007,46(47):13523-13532
Transition-metal ions (Cu2+ and Zn2+) play critical roles in the Abeta plaque formation. However, precise roles of the metal ions in the Abeta amyloidogenesis have been controversial. In this study, the molecular mechanism of the metal-induced Abeta oligomerization was investigated with extensive metal ion titration NMR experiments. Upon additions of the metal ions, the N-terminal region (1-16) of the Abeta (1-40) peptide was selectively perturbed. In particular, polar residues 4-8 and 13-15 were more strongly affected by the metal ions, suggesting that those regions may be the major binding sites of the metal ions. The NMR signal changes of the N-terminal region were dependent on the peptide concentrations (higher peptide concentrations resulted in stronger signal changes), suggesting that the metal ions facilitate the intermolecular contact between the Abeta peptides. The Abeta (1-40) peptides (>30 microM) were eventually oligomerized even at low temperature (3 degrees C), where the Abeta peptides are stable as monomeric forms without the metal ions. The real-time oligomerization process was monitored by 1H/15N HSQC NMR experiments, which provided the first residue-specific structural transition information. Hydrophobic residues 12-21 initially underwent conformational changes due to the intermolecular interactions. After the initial structural rearrangements, the C-terminal residues (32-40) readjusted their conformations presumably for effective oligomerization. Similar structural changes of the metal-free Abeta (1-40) peptides were also observed in the presence of the preformed oligomers, suggesting that the conformational transitions may be the general molecular mechanism of the Abeta (1-40) amyloidogenesis.  相似文献   

3.
Alzheimer's disease is a debilitating neurodegenerative disorder associated with the abnormal self-assembly of amyloid-beta (Abeta) peptides into fibrillar species. N-methylated peptides homologous to the central hydrophobic core of the Abeta peptide are potent inhibitors of this aggregation process. In this work, we use fully atomistic molecular dynamics simulations to study the interactions of the N-methylated peptide inhibitor Abeta16-20m (Ac-Lys(16)-(Me)Leu(17)-Val(18)-(Me)Phe(19)-Phe(20)-NH(2)) with a model protofilament consisting of Alzheimer Abeta16-22 peptides. Our simulations indicate that the inhibitor peptide can bind to the protofilament at four different sites: 1), at the edge of the protofilament; 2), on the exposed face of a protofilament layer; 3), between the protofilament layers; and 4), between the protofilament strands. The different binding scenarios suggest several mechanisms of fibrillogenesis inhibition: 1), fibril inhibition of longitudinal growth (in the direction of monomer deposition); 2), fibril inhibition of lateral growth (in the direction of protofilament assembly); and 3), fibril disassembly by strand removal and perturbation of the periodicity of the protofilament (disruption of fibril morphology). Our simulations suggest that the Abeta16-20m inhibitor can act on both prefibrillar species and mature fibers and that the specific mechanism of inhibition may depend on the structural nature of the Abeta aggregate. Disassembly of the fibril can be explained by a mechanism through which the inhibitor peptides bind to disaggregated or otherwise free Abeta16-22 peptides in solution, leading to a shift in the equilibrium from a fibrillar state to one dominated by inhibitor-bound Abeta16-22 peptides.  相似文献   

4.
Alpha(2)-macroglobulin (alpha(2)M) and its receptor, low density lipoprotein receptor-related protein (LRP), function together to facilitate the cellular uptake and degradation of beta-amyloid peptide (Abeta). In this study, we demonstrate that Abeta binds selectively to alpha(2)M that has been induced to undergo conformational change by reaction with methylamine. Denatured alpha(2)M subunits, which were immobilized on polyvinylidene difluoride membranes, bound Abeta, suggesting that alpha(2)M tertiary and quaternary structure are not necessary. To determine whether a specific sequence in alpha(2)M is responsible for Abeta binding, we prepared and analyzed defined alpha(2)M fragments and glutathione S-transferase-alpha(2)M peptide fusion proteins. A single sequence, centered at amino acids (aa) 1314-1365, was identified as the only major Abeta-binding site. Importantly, Abeta did not bind to the previously characterized growth factor-binding site (aa 718-734). Although the Abeta binding sequence is adjacent to the binding site for LRP, the results of experiments with mutated fusion proteins indicate that the two sites are distinct. Furthermore, a saturating concentration of Abeta did not inhibit LRP-mediated clearance of alpha(2)M-MA in mice. Using various methods, we determined that the K(D) for the interaction of Abeta with its binding site in the individual alpha(2)M subunit is 0.7-2.4 microm. The capacity of alpha(2)M to bind Abeta and deliver it to LRP may be greater than that predicted by the K(D), because each alpha(2)M subunit may bind Abeta and the bound Abeta may multimerize. These studies suggest a model in which alpha(2)M has three protein interaction sites with distinct specificities, mediating the interaction with Abeta, growth factors, and LRP.  相似文献   

5.
Recent studies show that quantitative and qualitative differences in amyloid beta (Abeta ) peptides may be implicated in the development of Alzheimer's disease. New evidence seems to support the existence of a dynamic equilibrium between Abeta peptide in the brain and peripheral blood circulation. The quantitation of Abeta in the blood may allow the development of the potential value of Abeta peptides as a biomarker in the development of Alzheimer's disease. In this communication, quantitation of Abeta peptides using high-performance liquid chromatography coupled with tandem mass spectrometry in a linear ion trap mode is presented. RP-HPLC was performed using a Waters Xterra MS C8 column (3.0 mm x 150 mm). Abeta(1-40) peptide was eluted using a gradient elution program. Eluate from the RP-HPLC column was split to both the UV detector and electrospray ionization MS source. The product ion scan was performed in a linear ion trap mode utilizing the transition of a multiply charged molecular ion of Abeta(1-40) to a singly charged product ion. The detection limit of 31.25 ng in column load using a 3.0-mm-diameter conventional C8 column was achieved. The Abeta(1-40) standard calibration curves show excellent linearity from 34 ng to 2500 ng Abeta(1-40) of column sample load. The product ion scan enhances sensitivity 10 times compared with the best previously achieved by a single-quadrupole instrument in the selective ion monitoring mode. Moreover, the product ion scan of Abeta(1-40) provides superior selectivity and specificity, which is very important in the quantitation of Abeta(1-40) in a complex biological matrix.  相似文献   

6.
Major constituents of the amyloid plaques found in the brain of Alzheimer's patients are the 39-43 residue beta-amyloid (Abeta) peptides. Extensive in vitro as well as in vivo biochemical studies have shown that the 40- and 42-residue Abeta peptides play major roles in the neurodegenerative pathology of Alzheimer's disease. Although the two Abeta peptides share common aggregation properties, the 42-residue peptide is more amyloidogenic and more strongly associated with amyloid pathology. Thus, characterizations of the two Abeta peptides are of critical importance in understanding the molecular mechanism of Abeta amyloid formation. In this report, we present combined CD and NMR studies of the monomeric states of the two peptides under both non-amyloidogenic (<5 degrees C) and amyloid-forming conditions (>5 degrees C) at physiological pH. Our CD studies of the Abeta peptides showed that initially unfolded Abeta peptides at low temperature (<5 degrees C) gradually underwent conformational changes to more beta-sheet-like monomeric intermediate states at stronger amyloidogenic conditions (higher temperatures). Detailed residue-specific information on the structural transition was obtained by using NMR spectroscopy. Residues in the N-terminal (3-12) and 20-22 regions underwent conformational changes to more extended structures at the stronger amyloidogenic conditions. Almost identical structural transitions of those residues were observed in the two Abeta peptides, suggesting a similar amyloidogenic intermediate for the two peptides. The 42-residue Abeta (1-42) peptide was, however, more significantly structured at the C-terminal region (39-42), which may lead to the different aggregation propensity of the two peptides.  相似文献   

7.
Multiple long molecular dynamics simulations are used to probe the oligomerization mechanism of Abeta(16-22) (KLVFFAE) peptides. The peptides, in the monomeric form, adopt either compact random-coil or extended beta strand-like structures. The assembly of the low-energy oligomers, in which the peptides form antiparallel beta sheets, occurs by multiple pathways with the formation of an obligatory alpha-helical intermediate. This observation and the experimental results on fibrillogenesis of Abeta(1-40) and Abeta(1-42) peptides suggest that the assembly mechanism (random coil --> alpha helix --> beta strand) is universal for this class of peptides. In Abeta(16-22) oligomers both interpeptide hydrophobic and electrostatic interactions are critical in the formation of the antiparallel beta sheet structure. Mutations of either hydrophobic or charged residues destabilize the oligomer, which implies that the 16-22 fragments of Arctic (E22G), Dutch (E22Q), and Italian (E22K) mutants are unlikely to form ordered fibrils.  相似文献   

8.
In order to clarify the basis of neuronal toxicity exerted by the shortest active peptides of amyloid beta-protein (Abeta), the toxic effects of Abeta(31-35) and Abeta(25-35) peptides on isolated rat brain mitochondria were investigated. The results show that exposure of isolated rat brain mitochondria to Abeta(31-35) and Abeta(25-35) peptides determines: (i) release of cytochrome c; (ii) mitochondrial swelling and (iii) a significant reduction in mitochondrial oxygen consumption. In contrast, the amplitude of these events resulted attenuated in isolated brain mitochondria exposed to the Abeta(31-35)Met35(OX) in which methionine-35 was oxidized to methionine sulfoxide. The Abeta peptide derivative with norleucine substituting Met-35, i.e., Abeta(31-35)Nle-35, had not effect on any of the biochemical parameters tested. We have further characterized the action of Abeta(31-35) and Abeta(25-35) peptides on neuronal cells. Taken together our result indicate that Abeta(31-35) and Abeta(25-35) peptides in non-aggregated form, i.e., predominantly monomeric, are strongly neurotoxic, having the ability to enter within the cells, determining mitochondrial damage with an evident trigger of apoptotic signals. Such a mechanism of toxicity seems to be dependent by the redox state of methionine-35.  相似文献   

9.
The amyloid beta peptides (Abeta) are the major components of the senile plaques characteristic of Alzheimer's disease. Abeta peptides are generated from the cleavage of amyloid precursor protein (APP) by beta- and gamma-secretases. Beta-secretase (BACE), a type-I transmembrane aspartyl protease, cleaves APP first to generate a 99-amino acid membrane-associated fragment (CT99) containing the N terminus of Abeta peptides. Gamma-secretase, a multi-protein complex, then cleaves within the transmembrane region of CT99 to generate the C termini of Abeta peptides. The production of Abeta peptides is, therefore, dependent on the activities of both BACE and gamma-secretase. The cleavage of APP by BACE is believed to be a prerequisite for gamma-secretase-mediated processing. In the present study, we provide evidence both in vitro and in cells that BACE-mediated cleavage between amino acid residues 34 and 35 (Abeta-34 site) in the Abeta region is dependent on gamma-secretase activity. In vitro, the Abeta-34 site is processed specifically by BACE1 and BACE2, but not by cathepsin D, a closely related aspartyl protease. Moreover, the cleavage of the Abeta-34 site by BACE1 or BACE2 occurred only when Abeta 1- 40 peptide, a gamma-secretase cleavage product, was used as substrate, not the non-cleaved CT99. In cells, overexpression of BACE1 or BACE2 dramatically increased the production of the Abeta 1-34 species. More importantly, the cellular production of Abeta 1-34 species induced by overexpression of BACE1 or BACE2 was blocked by a number of known gamma-secretase inhibitors in a concentration-dependent manner. These gamma-secretase inhibitors had no effect on enzymatic activity of BACE1 or BACE2 in vitro. Our data thus suggest that gamma-secretase cleavage of CT99 is a prerequisite for BACE-mediated processing at Abeta-34 site. Therefore, BACE and gamma-secretase activity can be mutually dependent.  相似文献   

10.
Extracellular accumulation of beta-amyloid peptide (Abeta) has been linked to the development of Alzheimer disease. The importance of intraneuronal Abeta has been recognized more recently. Although considerable evidence indicates that extracellular Abeta contributes to the intracellular pool of Abeta, the mechanisms involved in Abeta uptake by neurons are poorly understood. We examined the molecular mechanisms involved in Abeta-(1-42) internalization by primary neurons in the absence of apolipoprotein E. We demonstrated that Abeta-(1-42) is more efficiently internalized by axons than by cell bodies of sympathetic neurons, suggesting that Abeta-(1-42) uptake might be mediated by proteins enriched in the axons. Although the acetylcholine receptor alpha7nAChR, previously suggested to be involved in Abeta internalization, is enriched in axons, our results indicate that it does not mediate Abeta-(1-42) internalization. Moreover, receptors of the low density lipoprotein receptor family are not essential for Abeta-(1-42) uptake in the absence of apolipoprotein E because receptor-associated protein had no effect on Abeta uptake. By expressing the inactive dynamin mutant dynK44A and the clathrin hub we found that Abeta-(1-42) internalization is independent of clathrin but dependent on dynamin, which suggests an endocytic pathway involving caveolae/lipid rafts. Confocal microscopy studies showing that Abeta did not co-localize with the early endosome marker EEA1 further support a clathrin-independent mechanism. The lack of co-localization of Abeta with caveolin in intracellular vesicles and the normal uptake of Abeta by neurons that do not express caveolin indicate that Abeta does not require caveolin either. Instead partial co-localization of Abeta-(1-42) with cholera toxin subunit B and sensitivity to reduction of cellular cholesterol and sphingolipid levels suggest a caveolae-independent, raft-mediated mechanism. Understanding the molecular events involved in neuronal Abeta internalization might identify potential therapeutic targets for Alzheimer disease.  相似文献   

11.
A growing body of evidence supports an important role for oxidative stress in the pathogenesis of Alzheimer's disease. Recently, a number of papers have shown a synergistic neurotoxicity of amyloid beta peptide and cupric ions. We hypothesized that complexes of cupric ions with neurotoxic amyloid beta peptides (Abeta) can stimulate copper-mediated free radical formation. We found that neurotoxic Abeta (1-42), Abeta (1-40), and Abeta (25-35) stimulated copper-mediated oxidation of ascorbate, whereas nontoxic Abeta (40-1) did not. Formation of ascorbate free radical was significantly increased by Abeta (1-42) in the presence of ceruloplasmin. Once cupric ion is reduced to cuprous ion, it can be oxidized by oxygen to generate superoxide radical or it can react with hydrogen peroxide to form hydroxyl radical. Hydrogen peroxide greatly increased the oxidation of cyclic hydroxylamines and ascorbate by cupric-amyloid beta peptide complexes, implying redox cycling of copper ions. Using the spin-trapping technique, we have shown that toxic amyloid beta peptides led to a 4-fold increase in copper-mediated hydroxyl radical formation. We conclude that toxic Abeta peptides do indeed stimulate copper-mediated oxidation of ascorbate and generation of hydroxyl radicals. Therefore, cupric-amyloid beta peptide-stimulated free radical generation may be involved in the pathogenesis of Alzheimer's disease.  相似文献   

12.
Cerebral amyloid angiopathy (CAA) due to amyloid beta-protein (Abeta) is a key pathological feature of patients with Alzheimer's disease and hereditary cerebral hemorrhage with amyloidosis, Dutch-type (HCHWA-D). The CAA in these disorders is characterized by deposition of Abeta in the smooth muscle cells within the cerebral vessel wall. Recently, a new mutation in Abeta, E22K, was identified in several Italian families that, like HCHWA-D, is associated with CAA and hemorrhagic stroke. These two similar disorders, stemming from amino acid substitutions at position 22 of Abeta, implicate the importance of this site in the pathology of HCHWA. Previously we showed that HCHWA-D Abeta(1-40) containing the E22Q substitution induces robust pathologic responses in cultured human cerebrovascular smooth muscle cells (HCSM cells), including highly elevated levels of cell-associated Abeta precursor (AbetaPP) and cell death. In the present study, a series of E22 mutant Abeta(1-40) peptides were synthesized, and their pathogenic properties toward cultured HCSM cells were evaluated. Quantitative fluorescence analyses showed that mutant Abeta(1-40) peptides either containing a loss of charge (E22Q and E22A) or a change of charge (E22K) bind to the surface of HCSM cells and form amyloid fibrils. Similarly, this same group of E22 mutant Abeta(1-40) peptides caused enhanced pathologic responses in HCSM cells. In contrast, wild-type E22 or the charge-preserving E22D Abeta(1-40) peptides were devoid of any of these pathogenic properties. These data suggest that a change or loss of charge at position 22 of Abeta enhances the pathogenic effects of the peptide toward HCSM cells and may contribute to the pathogenesis of the phenotypically related HCHWA disorders.  相似文献   

13.
The detailed analysis of beta-amyloid (Abeta) peptides in human plasma is still hampered by the limited sensitivity of available mass spectrometric methods and the lack of appropiate ELISAs to measure Abeta peptides other than Abeta(1-38), Abeta(1-40), and Abeta(1-42). By combining high-yield Abeta immuno- precipitation (IP), IEF, and urea-based Abeta-SDS-PAGE-immunoblot, at least 30 Abeta-immuno-reactive spots were detected in human plasma samples as small as 1.6 mL. This approach clearly resolved Abeta peptides Abeta(1-40), Abeta(1-42), Abeta(1-37), Abeta(1-38), Abeta(1-39), the N-truncated Abeta(2-40), Abeta(2-42), and, for the first time, also Abeta(1-41). Relative quantification indicated that Abeta(1-40) and Abeta(1-42) accounted for less than 60% of the total amount of Abeta peptides in plasma. All other Abeta peptides appear to be either C-terminally or N-terminally truncated forms or as yet uncharacterized Abeta species which migrated as trains of spots with distinct pIs. The Abeta pattern found in cerebrospinal fluid (CSF) was substantially less complex. This sensitive method (2-D Abeta-WIB) might help clarifying the origin of distinct Abeta species from different tissues, cell types, or intracellular pools as well as their amyloidogenicity. It might further help identifying plasma Abeta species suitable as biomarkers for the diagnosis of Alzheimer's disease (AD).  相似文献   

14.
Chebaro Y  Derreumaux P 《Proteins》2009,75(2):442-452
Aggregation of the Abeta1-40/Abeta1-42 peptides is a key factor in Alzheimer's disease. Though the inhibitory effect of N-methylated Abeta16-22 (mAbeta16-22) peptides is well characterized in vitro, there is little information on how they disassemble full-length Abeta fibrils or block fibril formation. Here, we report coarse-grained implicit solvent molecular dynamics (MD) and replica exchange molecular dynamics (REMD) simulations on Abeta16-22 and mAbeta16-22 monomers, and then a preformed six-chain Abeta16-22 bilayer with either four copies of Abeta16-22 or four copies of mAbeta16-22. Our simulations show that the effect of N-methylation on mAbeta16-22 monomer is to reduce the density of compact forms. While 100 ns MD trajectories do not reveal any significant differences between the two ten-chain systems, the REMD simulations totaling 1 micros help understand the first steps of Abeta16-22 protofibril disassembly by N-methylated inhibitors. Notably, we find that mAbeta16-22 preferentially interacts with Abeta16-22 by blocking both beta-sheet extension and lateral association of layers, but also by intercalation of the inhibitors allowing sequestration of Abeta16-22 peptides. This third binding mode is particularly appealing for blocking Abeta fibrillogenesis.  相似文献   

15.
NMRsolution structures are reported for two mutants (K16E, K16F) of the soluble amyloid beta peptide Abeta(1-28). The structural effects of these mutations of a positively charged residue to anionic and hydrophobic residues at the alpha-secretase cleavage site (Lys16-Leu17) were examined in the membrane-simulating solvent aqueous SDS micelles. Overall the three-dimensional structures were similar to that for the native Abeta(1-28) sequence in that they contained an unstructured N-terminus and a helical C-terminus. These structural elements are similar to those seen in the corresponding regions of full-length Abeta peptides Abeta(1-40) and Abeta(1-42), showing that the shorter peptides are valid model systems. The K16E mutation, which might be expected to stabilize the macrodipole of the helix, slightly increased the helix length (residues 13-24) relative to the K16F mutation, which shortened the helix to between residues 16 and 24. The observed sequence-dependent control over conformation in this region provides an insight into possible conformational switching roles of mutations in the amyloid precursor protein from which Abeta peptides are derived. In addition, if conformational transitions from helix to random coil to sheet precede aggregation of Abeta peptides in vivo, as they do in vitro, the conformation-inducing effects of mutations at Lys16 may also influence aggregation and fibril formation.  相似文献   

16.
Human lumbar CSF patterns of Abeta peptides were analysed by urea-based beta-amyloid sodium dodecyl sulphate polyacrylamide gel electrophoresis with western immunoblot (Abeta-SDS-PAGE/immunoblot). A highly conserved pattern of carboxyterminally truncated Abeta1-37/38/39 was found in addition to Abeta1-40 and Abeta1-42. Remarkably, Abeta1-38 was present at a higher concentration than Abeta1-42, being the second prominent Abeta peptide species in CSF. Patients with Alzheimer's disease (AD, n = 12) and patients with chronic inflammatory CNS disease (CID, n = 10) were differentiated by unique CSF Abeta peptide patterns from patients with other neuropsychiatric diseases (OND, n = 37). This became evident only when we investigated the amount of Abeta peptides relative to their total Abeta peptide concentration (Abeta1-x%, fractional Abeta peptide pattern), which may reflect disease-specific gamma-secretase activities. Remarkably, patients with AD and CID shared elevated Abeta1-38% values, whereas otherwise the patterns were distinct, allowing separation of AD from CID or OND patients without overlap. The presence of one or two ApoE epsilon4 alleles resulted in an overall reduction of CSF Abeta peptides, which was pronounced for Abeta1-42. The severity of dementia was significantly correlated to the fractional Abeta peptide pattern but not to the absolute Abeta peptide concentrations.  相似文献   

17.
To investigate the molecular background of vestibular compensation, a model of lesion-induced plasticity, we used a microarray analysis to examine genes that show asymmetrical expression between the bilateral vestibular nucleus complexes (VNCs) 6 h following unilateral vestibular deafferentation (UVD). Asymmetrical gene expression was then validated by a real-time quantitative PCR. Among the 88 genes for which the ipsilateral (ipsi) : contralateral (contra) was > 1.35, the number of known genes was 33 (38%), and the number of expressed sequence tag (EST) sequences was 55 (62%). Among the 130 genes for which the contra : ipsi was > 1.35, the number of known genes was 55 (42%), and the number of EST sequences was 75 (58%). Changes in some of the genes were consistent with previous studies; however, we found several new genes which could be functionally related to the molecular basis of the electrophysiological asymmetry between the VNCs following UVD. Ipsi > contra genes included the GABA(A) receptor rho subunit, regulatory proteins of G protein signaling, calcium signaling related molecules such as the voltage-dependent calcium channel alpha2/delta subunit 1, calcineurin subunit Abeta and Ca(2+) pump. Contra > ipsi genes included the neuronal high affinity glutamate transporter, 5-hydroxytryptamine receptor 1D, mitogen-activated protein kinase 12 and ubiquitin carboxy-terminal hydrolase L1.  相似文献   

18.
In a previous article (Zbilut et al., Biophys J 2003;85:3544-3557), we demonstrated how an aggregation versus folding choice could be approached considering hydrophobicity distribution and charge. In this work, our aim is highlighting the mutual interaction of charge and hydrophobicity distribution in the aggregation process. Use was made of two different peptides, both derived from a transmembrane protein (amyloid precursor protein; APP), namely, Abeta(1-28) and Abeta(1-40). Abeta(1-28) has a much lower aggregation propensity than Abeta(1-40). The results obtained by means of molecular dynamics simulations show that, when submitted to the most "aggregation-prone" environment, corresponding to the isoelectric point and consequently to zero net charge, both peptides acquire their maximum flexibility, but Abeta(1-40) has a definitely higher conformational mobility than Abeta(1-28). The absence of a hydrophobic "tail," which is the most mobile part of the molecule in Abeta(1-40), is the element lacking in Abeta(1-28) for obtaining a "fully aggregating" phenotype. Our results suggest that conformational flexibility, determined by both hydrophobicity and charge effect, is the main mechanistic determinant of aggregation propensity.  相似文献   

19.
Alzheimer's disease (AD) is characterized by overproduction of beta amyloid peptides in the brain with progressive loss of neuronal cells. The 42-aa form of the beta amyloid peptide (Abeta(42)) is implied as a major causative factor, because it is toxic to neurons and elicits inflammatory responses in the brain by activating microglial cells. Despite the overproduction of Abeta(42), AD brain tissue also generates protective factor(s) that may antagonize the neurodestructive effect of Abeta(42). Humanin is a gene cloned from an apparently normal region of an AD brain and encodes a 24-aa peptide. Both secreted and synthetic Humanin peptides protect neuronal cells from damage by Abeta(42), and the effect of Humanin may involve putative cellular receptor(s). To elucidate the molecular identity of such receptor(s), we examined the activity of synthetic Humanin on various cells and found that Humanin induced chemotaxis of mononuclear phagocytes by using a human G protein-coupled formylpeptide receptor-like-1 (FPRL1) and its murine counterpart FPR2. Coincidentally, FPRL1 and FPR2 are also functional receptors used by Abeta(42) to chemoattract and activate phagocytic cells. Humanin reduced the aggregation and fibrillary formation by suppressing the effect of Abeta(42) on mononuclear phagocytes. In neuroblast cells, Humanin and Abeta(42) both activated FPRL1; however, only Abeta(42) caused apoptotic death of the cells, and its cytopathic effect was blocked by Humanin. We conclude that Humanin shares human FPRL1 and mouse FPR2 with Abeta(42) and suggest that Humanin may exert its neuroprotective effects by competitively inhibiting the access of FPRL1 to Abeta(42).  相似文献   

20.
Beta-amyloid peptides (Abeta) are the main protein components of neuritic plaques and are important in the pathogenesis of Alzheimer's disease. It is reported that Abeta itself is not toxic; however, it becomes toxic to neuronal cells once it has aggregated into amyloid fibrils by peptide-peptide interactions. In this study, to specify the molecular mechanism of aggregation, a novel fluorescence assay was designed. For this purpose, possible partial peptides (38 types of 5-mer) were synthesized on solid-phase. The molecular interactions were examined by a fluorescence probe possessing Lys-Leu-Val-Phe-Phe (KLVFF) as a molecular recognition site. KLVFF is known to be a minimum sequence for formation of the Abeta aggregate. A specific interaction was observed between labeled and immobilized KLVFF. It suggests that the aggregation of Abeta was controlled by the recognition of KLVFF itself by hydrophobic and electrostatic interactions.  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