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.
The aggregation of soluble, nontoxic amyloid beta (Abeta) peptide to beta-sheet containing fibrils is assumed to be a major step in the development of Alzheimer's disease. Interactions of Abeta with neuronal membranes could play a key role in the pathogenesis of the disease. Herein, we study the adsorption of synthetic Abeta peptide to DPPE and DMPE monolayers (dipalmitoyl- and dimyristoylphosphatidylethanolamine). Both lipids exhibit a condensed monolayer state at 20 degrees C and form a similar lattice. However, at low packing densities (at large area per molecule), the length of the acyl chains determines the phase behavior, therefore DPPE is fully condensed whereas DMPE exhibits a liquid-expanded state with a phase transition at approximately 5-6 mNm(-1). Adsorption of Abeta to DPPE and DMPE monolayers at low surface pressure leads to an increase of the surface pressure to approximately 17 mNm(-1). The same was observed during adsorption of the peptide to a pure air-water interface. Grazing incidence X-ray diffraction (GIXD) experiments show no influence of Abeta on the lipid structure. The adsorption kinetics of Abeta to a DMPE monolayer followed by IRRAS (infrared reflection absorption spectroscopy) reveals the phase transition of DMPE molecules from liquid-expanded to condensed states at the same surface pressure as for DMPE on pure water. These facts indicate no specific interactions of the peptide with either lipid. In addition, no adsorption or penetration of the peptide into the lipid monolayers was observed at surface pressures above 30 mNm(-1). IRRAS allows the measurement of the conformation and orientation of the peptide adsorbed to the air-water interface and to a lipid monolayer. In both cases, with lipids at surface pressures below 20 mNm(-1) and at the air-water interface, adsorbed Abeta has a beta-sheet conformation and these beta-sheets are oriented parallel to the interface.  相似文献   

2.
Modulation of the fibrillogenesis of amyloid peptide Abeta(1-40) with the cationic gemini surfactant hexamethylene-1,6-bis(dodecyldimethylammonium bromide) (C(12)C(6)C(12)Br(2)) has been studied. Both UV-vis and AFM results show that C(12)C(6)C(12)Br(2) monomers can promote the fibrillogenesis of Abeta(1-40) while its micelles inhibit this process. The electrostatic/hydrophobic force balance plays important roles in determining the Abeta(1-40) aggregation style and the secondary structures. When the surfactant positive charges are close to the Abeta(1-40) negative charges in number, the hydrophobic interaction is highly enhanced in the system. Both the nucleation rate and the lateral association between fibrils are greatly promoted. However, when the surfactant positive charges are in excess of the Abeta(1-40) negative charges, the electrostatic interaction is strengthened. In this case, the lateral association is inhibited and the alpha-helix to beta-sheet transition in the secondary structure is prevented. Simultaneously, another assembly pathway is induced to give the amorphous aggregates. Moreover, the size and surface roughness of the Abeta(1-40) aggregates also vary upon increasing C(12)C(6)C(12)Br(2) concentration.  相似文献   

3.
The aggregation of amyloid beta-peptide [Abeta(1-40)] into fibril is a key pathological process associated with Alzheimer's disease. The effect of cationic gemini surfactant hexamethylene-1,6-bis-(dodecyldimethylammonium bromide) [C(12)H(25)(CH(3))(2)N(CH(2))(6)N(CH(3))(2)C(12)H(25)]Br(2) (designated as C(12)C(6)C(12)Br(2)) and single-chain cationic surfactant dodecyltrimethylammonium bromide (DTAB) on the Alzheimer amyloid beta-peptide Abeta(1-40) aggregation behavior was studied by microcalorimetry, circular dichroism (CD), and atomic force microscopy (AFM) measurements at pH 7.4. Without addition of surfactant, 0.5 g/L Abeta(1-40) mainly exists in dimeric state. It is found that the addition of the monomers of C(12)C(6)C(12)Br(2) and DTAB may cause the rapid aggregation of Abeta(1-40) and the fibrillar structures are observed by CD spectra and the AFM images. Due to the repulsive interaction among the head groups of surfactants and the formation of a small hydrophobic cluster of surfactant molecules, the fibrillar structure is disrupted again as the surfactant monomer concentration is increased, whereas globular species are observed in the presence of micellar solution. Different from single-chain surfactant, C(12)C(6)C(12)Br(2) has a much stronger interaction with Abeta(1-40) to generate larger endothermic energy at much lower surfactant concentration and has much stronger ability to induce the aggregation of Abeta(1-40).  相似文献   

4.
The aggregation of amyloid beta-peptide (Abeta(1-40)) into fibrils is a key pathological process associated with Alzheimer's disease. This work has investigated the micellization process of biosurfactant surfactin and its effect on the aggregation behavior of Abeta(1-40). The results show that surfactin has strong self-assembly ability to form micelles and the micelles tend to form larger aggregates. Surfactin adopts a beta-turn conformation at low micelle concentration but a beta-sheet conformation at high micelle concentration. The effect of surfactin on the Abeta(1-40) aggregation behavior exhibits a strong concentration-dependent fashion. Below the critical micelle concentration of surfactin, the electrostatic binding of surfactin monomers on Abeta(1-40) causes Abeta(1-40) molecules to unfold. Assisted by the hydrophobic interaction among surfactin monomers on the Abeta(1-40) chain, the conformation of Abeta(1-40) transfers to the beta-sheet structure, which promotes the formation of fibrils. At low surfactin micelle concentration, besides the electrostatic force and hydrophobic interaction, hydrogen bonds formed between surfactin micelles and adjacent Abeta(1-40) peptide chains may promote the ordered organization of these Abeta(1-40) peptide chains, thus leading to the formation of beta-sheets and fibrils to a great extent. At high surfactin micelle concentration, the separating of Abeta(1-40) chains by the excessive surfactin micelles and the aggregation of the complexes of Abeta(1-40) with surfactin micelles inhibit the formation of beta-sheets and fibrils.  相似文献   

5.
Results from replica-exchange and regular room temperature molecular dynamics simulations of the Alzheimer's beta amyloid (Abeta(1-39)) monomer in an implicit solvent are reported. Our data indicate that at room temperature, the monomer assumes random-coil and soluble conformations. No beta content is observed which therefore seems to be a product of oligomerization and aggregation of monomers.  相似文献   

6.
The pathogenesis of Alzheimer's disease is characterized by the aggregation and fibrillation of the 40-residue A beta(1-40) and 42-residue A beta(1-42) peptides into amyloid plaques. The structural changes associated with the conversion of monomeric A beta peptide building blocks into multimeric fibrillar beta-strand aggregates remain unknown. Recently, we established that oxidation of the methionine-35 side chain to the sulfoxide (Met35(red) --> Met35(ox)) significantly impedes the rate of aggregation and fibrillation of the A beta peptide. To explore this effect at greater resolution, we carefully compared the (1)H, (15)N, and (13)C NMR chemical shifts of four A beta peptides that had the Met35 reduced or oxidized (A beta(1-40)Met35(red), A beta(1-40)Met35(ox), A beta(1-42)Met35(red), and A beta(1-42)Met35(ox)). With the use of a special disaggregation protocol, the highly aggregation prone A beta peptides could be studied at higher, millimolar concentrations (as required by NMR) in aqueous solution at neutral pH, remaining largely monomeric at 5 degrees C as determined by sedimentation equilibrium studies. The NOE, amide-NH temperature coefficients, and chemical shift indices of the (1)H alpha, (13)C alpha, and (13)C beta established that the four peptides are largely random, extended chain structures, with the Met35(ox) reducing the propensity for beta-strand structure at two hydrophobic regions (Leu17-Ala21 and Ile31-Val36), and turn- or bendlike structures at Asp7-Glu11 and Phe20-Ser26. Additional NMR studies monitoring changes that occur during aging at 37 degrees C established that, along with a gradual loss of signal/noise, the Met35(ox) significantly hindered upfield chemical shift movements of the 2H NMR signals for the His6, His13, and His14 side chains. Taken together, the present NMR studies demonstrate that the Met35(red) --> Met35(ox) conversion prevents aggregation by reducing both hydrophobic and electrostatic association and that the A beta(1-40)Met35(red), A beta(1-40)Met35(ox), A beta(1-42)Met35(red), and A beta(1-42)Met35(ox) peptides may associate differently, through specific, sharp changes in structure during the initial stages of aggregation.  相似文献   

7.
The conformational change of the 39-43 residues of the amyloid beta-peptide (Abeta) toward a beta-sheet enriched state promotes self-aggregation of the peptide molecules and constitutes the major peptide component of the amyloid plaques in Alzheimer patients. The crucial question behind the self-aggregation of Abeta is related to the different pathways the peptide may take after cleavage from the amyloid precursor proteins at cellular membranes. This work is aiming at determining the conformation of the Abeta (1-40) adsorbed on hydrophobic Teflon and hydrophilic silica particles, as model sorbent surfaces mimicking the apolar transmembrane environment and the polar, charged membrane surface, respectively. The mechanism by which the Abeta interacts with solid surfaces strongly depends on the hydrophobic/hydrophilic character of the particles. Hydrophobic and electrostatic interactions contribute differently in each case, causing a completely different conformational change of the adsorbed molecules on the two surfaces. When hydrophobic interactions between the peptide and the sorbent prevail, the adsorbed Abeta (1-40) mainly adopts an alpha-helix conformation due to H-bonding in the apolar part of the peptide that is oriented towards the surface. On the other hand, when the peptide adsorbs by electrostatic interactions beta-sheet formation is promoted due to intermolecular association between the apolar parts of the adsorbed peptide. Irrespective of the characteristics of the solid sorbent, crowding the surface results in intermolecular association between adsorbed molecules leading to a strong aggregation tendency of the Abeta (1-40). [Diagram: see text] CD spectra of Abeta (1-40) at pH 7: A) in solution ([Abeta]=0.2 mg.ml(-1)) freshly prepared (line) and after overnight incubation (symbols);B) on Teflon (Gamma=0.5 mg.m(-2)).  相似文献   

8.
In this study, all-atom 50 ns molecular dynamics simulations are performed on the full-length amyloid beta (Abeta) monomers (WT-Abeta(1-40) and WT-Abeta(1-42)) and their oxidized forms (Met35(O)-Abeta(1-40) and Met35(O)-Abeta(1-42)) in aqueous solution. The effects of the oxidation state of Met35 and the presence of dipeptide (Ile41-Ala42) on the secondary structures of the three distinct regions (the central hydrophobic core region 17-21 (LVFFA), the loop 23-28 (DVGSNK), and the second hydrophobic domain 29-35 (GAIIGLM)) of all monomers have been analyzed in detail, and results are compared with the available experimental information. Our simulations indicate that the WT-Abeta(1-40) monomer adopts an overall beta-hairpin-like structure, which is promoted by the turn region (24-27). This turn region is stabilized through salt-bridge formation between the Asp23 and Lys28 residues. In contrast, the overall structure of the oxidized (Met35(O)-Abeta(1-40)) monomer can be divided into three well-defined bend regions separated by coil segments. These structural differences may be critical for the measured decrease in the rate of aggregation of Met35(O)-Abeta(1-40) peptide. In the WT-Abeta(1-42) monomer, in comparison to the WT-Abeta(1-40), the Asp23-Lys28 salt bridge is absent, and consequently, the turn in the middle (24-27) region has a smaller curvature. The observed difference in the aggregation rates of these two peptides may be related to the opening of the turn (24-27) stabilized by the Asp23-Lys28 salt bridge. For WT-Abeta(1-42), in the absence of this salt bridge, the unfolding and aggregation events may be more favorable than for WT-Abeta(1-40).  相似文献   

9.
The aggregation of full-length (residues 1-40) amyloid beta-peptide (A beta) and fragments corresponding to residues 1-23 and 24-40 was studied by electrospray mass spectrometry, using gramicidin as a non-aggregating reference. Following a lag period, A beta(1-40) at 140 microM concentration aggregates with apparent first-order kinetics. Under acidic conditions A beta(1-40) undergoes spontaneous cleavage between Asp23-Val24 and to a lesser extent also at two other Asp-X motifs. Incubation in acidic H(2)18O showed incorporation of 18O in fragment A beta(1-23), confirming that the Asp23-Val24 peptide bond had been hydrolyzed. Incubation of synthetic A beta(1-23) and A beta(24-40) peptides with A beta(1-40) showed that A beta(24-40) remained in solution for several months, that A beta(1-23) partly disappeared from solution, whereas A beta(1-40) completely disappeared. Further, treatment of sedimentable aggregates formed after co-incubation of the three peptides with hexafluoro-2-propanol or formic acid recovered the intensity of A beta(1-40). These data support previous studies showing that the region of A beta encompassing residues 16-24 is necessary for aggregation into amyloid fibrils.  相似文献   

10.
A pathogenetic feature of Alzhemier disease is the aggregation of monomeric beta-amyloid proteins (Abeta) to form oligomers. Usually these oligomers of long peptides aggregate on time scales of microseconds or longer, making computational studies using atomistic molecular dynamics models prohibitively expensive and making it essential to develop computational models that are cheaper and at the same time faithful to physical features of the process. We benchmark the ability of our implicit solvent model to describe equilibrium and dynamic properties of monomeric Abeta(10-35) using all-atom Langevin dynamics (LD) simulations, since Alphabeta(10-35) is the only fragment whose monomeric properties have been measured. The accuracy of the implicit solvent model is tested by comparing its predictions with experiment and with those from a new explicit water MD simulation, (performed using CHARMM and the TIP3P water model) which is approximately 200 times slower than the implicit water simulations. The dependence on force field is investigated by running multiple trajectories for Alphabeta(10-35) using the CHARMM, OPLS-aal, and GS-AMBER94 force fields, whereas the convergence to equilibrium is tested for each force field by beginning separate trajectories from the native NMR structure, a completely stretched structure, and from unfolded initial structures. The NMR order parameter, S2, is computed for each trajectory and is compared with experimental data to assess the best choice for treating aggregates of Alphabeta. The computed order parameters vary significantly with force field. Explicit and implicit solvent simulations using the CHARMM force fields display excellent agreement with each other and once again support the accuracy of the implicit solvent model. Alphabeta(10-35) exhibits great flexibility, consistent with experiment data for the monomer in solution, while maintaining a general strand-loop-strand motif with a solvent-exposed hydrophobic patch that is believed to be important for aggregation. Finally, equilibration of the peptide structure requires an implicit solvent LD simulation as long as 30 ns.  相似文献   

11.
Beta-azidoalanine dipeptide 1 was synthesized, and its azido stretching vibration in H2O and dimethyl sulfoxide (DMSO) was studied by using Fourier transform (FT) IR spectroscopy. The dipole strength of the azido stretch mode is found to be about 19 and 5 times larger than those of the CN and SCN stretch modes, respectively, which have been used as local environmental IR sensors. The azido stretch band in H2O is blue-shifted by about 14 cm(-1) in comparison to that in DMSO, indicative of its sensitivity to the electrostatic environment. To test the utility of beta-azidoalanine as an IR probe of the local electrostatic environment in proteins, azidopeptide 4 was prepared by its incorporation into Abeta(16-22) peptide of the Alzheimer's disease amyloid beta-protein at position Ala21. The amide I IR spectrum of 4 in D2O suggests that the azidopeptide thus modified forms in-register beta-sheets in aggregates as observed for normal Abeta(16-22). The azido peak frequency of 4 in aggregates is almost identical to that in DMSO, indicating that the azido group is not exposed to water but to the hydrophobic environment. We believe that beta-azidoalanine will be used as an effective IR probe for providing site-specific information about the local electrostatic environments of proteins.  相似文献   

12.
The possibility to monitor, in solution, the steps of beta-amyloid (Abeta) nucleation and therefore to describe this dynamic process by using capillary electrophoresis and under optimized experimental conditions is described. Striking differences in the electrophoretic patterns of Abeta 1-42 and Abeta 1-40 over time are here shown, and different aggregation states are elucidated, which reflect the very diverse oligomerization behavior of two very similar peptides. The isolation of one aggregated species of high molecular weight by ultracentrifugation allowed us to assess its role as toxic oligomer. The perturbation of the existing equilibrium among the identified species by the addition of small molecules can in principle interfere with the aggregation process of the peptides and ultimately prevent the plaque formation in vitro.  相似文献   

13.
The primary pathological characteristic of Alzheimer's disease is the presence in the brain of self-assembled beta amyloid (Abeta) protein fibrils, consisting of 35-43 amino acid residues. The toxicity of the aggregated protein structures has previously been proposed to be related to the interaction of Abeta fibrils with neuronal membranes (phospholipid bilayers). Here, surfaces consisting of self-assembled alkanethiol monolayers with different end groups--supported on Au--are used to test the effect of surface chemistry on the structure and morphology of aggregates formed from an active fragment (Abeta10-35) of the Abeta peptide. The influence of monolayer nature (end group) on the aggregation of Abeta10-35 was examined using reflection-absorption infrared spectroscopy (RAIRS) and scanning force microscopy (SFM). Evaluation of the SFM and RAIRS data reveals the presence of Abeta10-35 protein on the various monolayer surfaces, with the surface protein possessing predominantly beta-sheet and random-coil conformations. Time-dependent studies of the extent of Abeta10-35 aggregation and deposition on the various surfaces and the effect of the monolayers on seeding of Abeta10-35 aggregates in solution are also discussed.  相似文献   

14.
It is generally postulated that amyloid-beta-peptides play a central role in the progressive neurodegeneration observed in Alzheimer's disease. Important pathological properties of these peptides, such as neurotoxicity and resistance to proteolytic degradation, depend on the ability of amyloid-beta-peptides to form beta-sheet structures and/or amyloid fibrils. Amyloid-beta-peptides are known to aggregate spontaneously in vitro with the formation of amyloid fibrils. The intervention on the amyloid-beta-peptides aggregation process can be envisaged as an approach to stopping or slowing the progression of Alzheimer's disease. In the last few years a number of small molecules have been reported to interfere with the in vitro aggregation of amyloid-beta-peptides. Melatonin, a hormone recently found to protect neurons against amyloid-beta-peptide toxicity, interacts with amyloid-beta-peptide (1-40) and amyloid-beta-peptide (1-42) and inhibits the progressive formation of beta-sheet and/or amyloid fibrils. These interactions between melatonin and the amyloid peptides have been demonstrated by circular dichroism (CD) and electron microscopy for amyloid-beta-peptide (1-40) and amyloid-beta-peptide (1-42) and by nuclear magnetic resonance (NMR) spectroscopy for amyloid-beta-peptide (1-40). Our electrospray ionization mass spectrometric (ESI-MS) studies also proved that there is a hydrophobic interaction between amyloid-beta-peptide (1-40) and melatonin and the proteolytic investigations suggested that the interaction took place on the 29-40 amyloid-beta-peptide segment. The wide-ranging application of these results would provide further information and help in biological research.  相似文献   

15.
We present the first electrochemical detection, characterization, and kinetic study of the aggregation of Alzheimer's disease (AD) amyloid beta peptides (Abeta-40, Abeta-42) using three different voltammetric techniques at a glassy carbon electrode (GCE). This method is based on detecting changes in the oxidation signal of tyrosine (Tyr) residue. As the peptides aggregate, there are structure conformational changes, which affect the degree of exposure of Tyr to the molecular surface of the peptides. The results show significant differences in the aggregation process between the two peptides, and these correlate highly with established techniques. The method is rapid and label-free, and the principle can be universally applied to other protein aggregation studies related to diseases, such as Huntington's, Parkinson's, and Creutzfeldt Jacob (CJD). This method could also be explored in screening for the effectiveness of AD therapies.  相似文献   

16.
Aβ (31–35) peptide and control peptides as well as full length Aβ (1–40) and Aβ (1–42) peptides were labelled with luminescent CdSe/ZnS quantum dots (QDs) to observe the morphology of amyloid fibers. A comparison was made between QDs and an organic dye, namely Dansyl group, which showed that the QDs present a much better contrast for imaging than the organic dye.  相似文献   

17.
Amyloid beta-peptide (Abeta) plays a critical role in Alzheimer's disease (AD). The monomeric state of Abeta can self-assemble into oligomers, protofibrils, and amyloid fibrils. Since the fibrils and soluble oligomers are believed to be responsible for AD, the construction of molecules capable of capturing these species could prove valuable as a means of detecting these potentially toxic species and of providing information pertinent for designing drugs effective against AD. To this aim, we have designed short peptides with various hydrophobicities based on the sequence of Abeta14-23, which is a critical region for amyloid fibril formation. The binding of the designed peptides to Abeta and the amplification of the formation of peptide amyloid-like fibrils coassembled with Abeta are elucidated. A fluorescence assay utilizing thioflavin T, known to bind specifically to amyloid fibrils, revealed that two designed peptides (LF and VF, with the leucine and valine residues, respectively, in the hydrophobic core region) could form amyloid-like fibrils effectively by using mature Abeta1-42 fibrils as nuclei. Peptide LF also coassembled with soluble Abeta oligomers into peptide fibrils. Various analyses, including immunostaining with gold nanoparticles, enzyme-linked immunosorbent assays, and size-exclusion chromatography, confirmed that the LF and VF peptides formed amyloid-like fibrils by capturing and incorporating Abeta1-42 aggregates into their peptide fibrils. In this system, small amounts of mature Abeta1-42 fibrils or soluble oligomers could be transformed into peptide fibrils and detected by amplifying the amyloid-like fibrils with the designed peptides.  相似文献   

18.
19.
Congo-Red nonreactive beta-amyloid (1-42) exhibits in gel electrophoresis (pH 8.82, 0.01 M ionic strength, 2 degree C) a surface charge density larger than that of the corresponding peptide of length (1-40), and a size indistinguishable from that of (1-40).  相似文献   

20.
Recent studies have reported that the cholinergic anti-inflammatory pathway regulates peripheral inflammatory responses via alpha7 nicotinic acetylcholine receptors (alpha7 nAChRs) and that acetylcholine and nicotine regulate the expression of proinflammatory mediators such as TNF-alpha and prostaglandin E2 in microglial cultures. In a previous study we showed that ATP released by beta-amyloid-stimulated microglia induced reactive oxygen species (ROS) production, in a process involving the P2X(7) receptor (P2X(7)R), in an autocrine fashion. These observations led us to investigate whether stimulation by nicotine could regulate fibrillar beta amyloid peptide (1-42) (fAbeta1-42)-induced ROS production by modulating ATP efflux-mediated Ca(2+) influx through P2X(7)R. Nicotine inhibited ROS generation in fAbeta(1-42)-stimulated microglial cells, and this inhibition was blocked by mecamylamine, a non-selective nAChR antagonist, and a-bungarotoxin, a selective alpha7 nAChR antagonist. Nicotine inhibited NADPH oxidase activation and completely blocked Ca(2+) influx in fAbeta(1-42)-stimulated microglia. Moreover, ATP release from fAbeta(1-42)-stimulated microglia was significantly suppressed by nicotine treatment. In contrast, nicotine did not inhibit 2',3'-O-(4-benzoyl)-benzoyl ATP (BzATP)-induced Ca(2+) influx, but inhibited ROS generation in BzATP-stimulated microglia, indicating an inhibitory effect of nicotine on a signaling process downstream of P2X(7)R. Taken together, these results suggest that the inhibitory effect of nicotine on ROS production in fAbeta1-42-stimulated microglia is mediated by indirect blockage of ATP release and by directly altering the signaling process downstream from P2X(7)R.  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