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.
Secondary structure of three amyloid b-peptides [A beta(1-28), A beta(1-40) and A beta(1-42)] in the solid state was respectively determined by Fourier transform infrared (FT-IR) microspectroscopy. Their thermal-dependent structural transformation were also investigated by FT-IR microspectroscopy equipped with a thermal analyzer. The present result demonstrates that the solid-state A beta(1-28), A beta(1-40) and A beta(1-42) peptides showed a significant IR spectral difference in the amide I and II bands. The secondary conformation of A beta(1-28) peptide was the combination of major beta-sheet and minor alpha-helix with little random coil structures, but A beta(1-40) peptide showed the co-existence of major beta-sheet and minor random coil with little alpha-helix structures. A beta(1-42) peptide mainly consisted of the predominant b-sheet structure. Although the intact A beta(1-28), A beta(1-40) or A beta(1-42) peptide exhibits a different secondary structure, a similar beta-conformation may form after thermal treatment. A thermal-dependent transition was found for solid A beta(1-28) and A beta(1-40) peptides near 40 degrees C and 45 degrees C, respectively. There was no transition temperature for solid A beta(1-42) peptide, however, due to only a very little level of alpha-helix and random coil structure containing in the solid A beta(1-42) peptide. The thermal denaturation plays an important role in the structural transformation from alpha-helix/random coil to beta-sheet.  相似文献   

2.
The body of evidence indicating that oligomers of amyloid beta(1-42) (Abeta(1-42)) produce toxicity to neurons, together with our demonstration that prostaglandin H(2) (PGH(2)) oligomerizes amyloid beta(1-42), led to the examination of the neurotoxicity of amyloid beta(1-42) treated with PGH(2). The neurotoxic effects of Abeta(1-42) incubated with PGH(2) was examined in primary cultures of cerebral neurons of mice, monitoring the reduction of 3-(4,5-dimethylthiazole-2-yl)-2,5-diphenyltetrazolium bromide (MTT) as an indicator of cell toxicity. Whereas Abeta(1-42) itself, incubated for 24 h, has little or no effect on MTT reduction, Abeta(1-42) 24 h after exposure to PGH(2) produced a marked inhibition of MTT reduction, comparable with the inhibition resulting from Abeta(1-42) that has been oligomerized by incubation for 6 days. Similar results were obtained when Abeta(1-42) was incubated with levuglandin E(2) (LGE(2)), a reactive aldehyde formed by spontaneous rearrangement of PGH(2). The oligomers formed from reaction of Abeta(1-42) with LGE(2) exhibit immunochemical similarity with amyloid-derived diffusible ligands (ADDLs), as determined by analysis of the products of reaction of Abeta(1-42) with LGE(2) using western blotting with an antibody that is selective for ADDLs.  相似文献   

3.
The Fmoc solid phase synthesis of A beta(1-40), a strongly aggregating peptide found in Alzheimer's disease brain, was performed using 2-hydroxy-4-methoxybenzyl (Hmb) backbone amide protection. Hmb-Gly residues were incorporated using N(alpha)-Fmoc-Hmb-Gly-OH rather than N,O-bisFmoc-Hmb-Gly-OPfp. Amino acid acylation of the sterically hindered Hmb-amino acids was monitored using 'semi-on-line' MALDI-TOF-MS in a novel application of this technique which significantly simplified the successful incorporation of these residues. Standard coupling conditions in N,N-dimethylformamide (DMF) were used throughout the synthesis. Comparative structural studies of acetyl-Hmb-protected and native A beta(1-40) were performed to investigate the structural basis of Hmb-mediated disaggregation. The incorporation of backbone amide protection was observed by circular dichroism spectroscopy and gel electrophoresis to strongly affect the solution structure of A beta(1-40). Despite the reported structure-breaking activity of Hmb groups, penta(acetyl-Hmb)A beta(1-40) was found to adopt both alpha-helix and intermolecular beta-sheet conformations. In 100% TFE a mixed alpha-helix/random coil structure was formed by the protected peptide indicating reduced alpha-helical propensity relative to A beta(1-40). The protected peptide formed beta-sheet structures in aqueous buffer. Gel electrophoresis indicated that, unlike native A beta(1-40), penta(acetyl-Hmb)A beta(1-40) did not form large aggregate species.  相似文献   

4.
NMR studies of amyloid beta-peptides (A beta) in aqueous solution provide a novel way in which to characterize the apparent Alzheimer's disease-related conformational polymorphism of A beta. In the aqueous medium, neither of the polypeptides A beta(1-40)(ox) or A beta(1-42)(ox) (both of which contain a methionine sulfoxide at position 35) is folded into a globular structure, but they both deviate from random coil behavior by local conformational preferences of several short segments along the amino-acid sequence. Differences between the solution structures of A beta(1-40)(ox) and A beta(1-42)(ox) are indicated only by decreased flexibility of the region from about residue 32 to the C-terminus in A beta(1-42)(ox) when compared to A beta(1-40)(ox). The lack of the observation of more extensive conformational differences between the two molecules is intriguing, considering that A beta(1-42)(ox) in aqueous solution has much higher plaque-competence than A beta(1-40)(ox).  相似文献   

5.
Thermodynamics of A beta(1-40) amyloid fibril elongation   总被引:1,自引:0,他引:1  
  相似文献   

6.
The venerable fluorescent probe of protein hydrophobic regions, 4,4(')-dianilino-1,1(')-binaphthyl-5,5(')-disulfonate (bis-ANS), unexpectedly increases in fluorescence with soluble beta(1-40) in acidic buffer solutions but reacts weakly with amyloid fibrils while other hydrophobic probes react with the fibrils. CD analysis correlates reaction with the probe with random coil/mixed conformations and alpha-helical forms of beta(1-40) in buffer solutions but less so with soluble beta-sheet forms or amyloid fibrils. The kinetics of the fluoroalcohol-induced interconversion of conformers can be followed by changes in bis-ANS fluorescence. Formation of the beta-sheet form in aqueous buffer is limited by a slow component (minutes) while fluoroalcohol-promoted changes between beta-sheet and alpha-helix occur over seconds. Variants of beta(1-40) such as beta(1-42) or the Dutch E22Q mutation of beta(1-40) and fragments beta(1-28), beta(12-28), beta(10-20 amide), and beta(10-35 amide) react with bis-ANS under conditions that do not support fibril formation. Primary amino acid sequence is important as beta(1-11) does not cause bis-ANS fluorescence while beta(1-16) does, but hydrophobicity is not as beta(25-35) and beta(15-20 amide) are unreactive. bis-ANS is a useful biophysical tool for characterizing particular, but not all, soluble Abeta conformations distinct from the fibrillar form of amyloid peptides detected by Thioflavin T.  相似文献   

7.
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   

8.
The abnormal aggregation and deposition of amyloid β protein (Aβ) on neuronal cells are critical to the onset of Alzheimer's disease. The entity (oligomers or fibrils) of toxic Aβ species responsible for the pathogenesis of the disease has been controversial. We have reported that the Aβ aggregates on ganglioside-rich domains of neuronal PC12 cells as well as in raft-like model membranes. Here, we identified toxic Aβ(1-40) aggregates formed with GM1-ganglioside-containing membranes. Aβ(1-40) was incubated with raft-like liposomes composed of GM1/cholesterol/sphingomyelin at 1:2:2 and 37 °C. After a lag period, toxic amyloid fibrils with a width of 12 nm were formed and subsequently laterally assembled with slight changes in their secondary structure as confirmed by viability assay, thioflavin-T fluorescence, circular dichroism, and transmission electron microscopy. In striking contrast, Aβ fibrils formed without membranes were thinner (6.7 nm) and much less toxic because of weaker binding to cell membranes and a smaller surface hydrophobicity. This study suggests that toxic Aβ(1-40) species formed on membranes are not soluble oligomers but amyloid fibrils and that Aβ(1-40) fibrils exhibit polymorphisms.  相似文献   

9.
LeVine H 《Biochemistry》2005,44(48):15937-15943
K114, (trans,trans)-1-bromo-2,5-bis(4-hydroxystyryl)benzene, is a fluorescent Congo Red analogue that binds tightly to amyloid fibrils, but not the monomeric proteins, with a concomitant enhancement in fluorescence. The mechanism for the low aqueous fluorescence and the subsequent enhancement by A beta(1-40) fibrils was investigated by fluorescence spectroscopy and binding analysis. K114's unusually low buffer fluorescence is due to self-quenching in sedimentable aggregates or micelles which upon interacting with amyloid fibrils undergo an enhancement in fluorescence intensity and shifts in the excitation and emission spectra. These spectral changes are suggestive of a stabilization of the phenolate anion, perhaps by hydrogen bonding, rather than an increase in the microenvironment dielectric constant or dye immobilization. 1,4-Bis(4-aminophenylethenyl)-2-methoxybenzene, which lacks the phenol moiety, and X-34, which contains a stabilized phenol (pK approximately 13.4), do not display the phenolate anion fluorescence in the presence of fibrils. The apparent affinity of K114 for fibril binding is 20-30 nM with a stoichiometry of 2.2 mol of K114/mol of A beta(1-40) monomer. Competition studies indicate that K114 and Congo Red share a site, but K114 does not bind to sites on A beta(1-40) fibrils for neutral benzothiazole (BTA-1), cationic thioflavin T, or the hydrophobic (S)-naproxen and (R)-ibuprofen molecules. Comparison of benzothiazole binding stoichiometry which has been suggested to reflect disease-relevant amyloid structures to that of Congo Red analogues which reflect total fibril content may be useful in defining biologically pertinent conformational forms of amyloid.  相似文献   

10.
The human beta-secretase enzyme, BACE1, mediates a critical step in the production of A beta(40) and A beta(42) peptides which are responsible for the severe neuronal cell death and insoluble amyloid plaques of Alzheimer's disease (AD). Several lines of evidence suggest that potent BACE1 inhibitors represent an attractive A beta-lowering strategy for AD. We designed a simple homogeneous time-resolved fluorescence (HTRF) assay which utilizes the fluorescence resonance energy transfer (FRET) pair europium and allophycocyanin for measuring BACE1 enzymatic activity in a high-throughput manner. Robust FRET was observed when an 18-amino-acid APP Swedish-synthetic peptide that was N-terminally labeled with europium cryptate and C-terminally biotinylated was incubated with streptavidin-coupled cross-linked allophycocyanin (SA-XL665). Purified BACE1 enzyme caused a time- and concentration-dependent linear change in FRET at low nanomolar enzyme concentrations. This assay was used to compare the autoprocessed "mature" BACE1 enzyme (sautoBACe1) and the soluble proBACE1 for activity and inhibition by selected peptidic BACE inhibitors. sautoBACE1 displayed only a modest increase in activity compared to sproBACE1 and this activity was uninhibited by the BACE1 prodomain peptide. Interestingly, the BACE1 prodomain peptide was able to partially inhibit sproBACE1 activity. IC(50s) for a P10-P4' statine BACE1 inhibitor, OM99-2, and OM-003 determined using the HTRF assay were in good agreement with those reported in the literature. The primary advantages of the HTRF-formatted BACE1 protease assay include appropriate reflection of native BACE1 activity, high sensitivity, low variability, and intrinsic quench correction afforded by ratiometric measurements made between EuK and SA-XL665 fluorophores.  相似文献   

11.
Sheikh AM  Nagai A 《The FEBS journal》2011,278(4):634-642
Phospholipids are known to influence fibril formation of amyloid beta (Aβ) peptide. Here, we show that lysophosphatidylcholine (LPC), a polar phospholipid, enhances Aβ(1-42) fibril formation, by decreasing the lag time and the critical peptide concentration required for fibril formation, and increasing the fibril elongation rate. Conversely, LPC did not have an enhancing effect on Aβ(1-40) fibril formation, and appeared to be inhibitory. Tyrosine fluorescence spectroscopy showed that LPC altered the fluorescence spectra of Aβ(1-40) and Aβ(1-42) in opposite ways. Further, 8-anilino-1-naphthalene sulfonic acid fluorescence spectroscopy showed that LPC significantly increased the hydrophobicity of Aβ(1-42), but not of Aβ(1-40). Tris-tricine gradient SDS/PAGE revealed that LPC increased the formation of higher-molecular-weight species of Aβ(1-42), including trimers and tetramers. LPC had no such effect on Aβ(1-40), and thus may specifically influence the oligomerization and nucleation processes of Aβ(1-42) in a manner dependent on its native structure. Dot-blot assays confirmed that LPC induced Aβ(1-42) oligomer formation at an early time point. Thus our results indicate that LPC specifically enhances the formation of Aβ(1-42) fibrils, the main component of senile plaques in Alzheimer's disease patients, and may be involved in Alzheimer's disease pathology.  相似文献   

12.
An N-terminal truncated isoform of the amyloid beta-peptide (A beta) that begins with a pyroglutamate (pE) residue at position 3 [A beta3(pE)-42] is the predominant isoform found in senile plaques. Based upon previous in vitro studies regarding A beta N-terminal truncated isoforms, it has been hypothesized that A beta3(pE)-x isoforms may aggregate more rapidly and become more toxic than corresponding Abeta1-x peptides. However, the toxicity and aggregation properties of A beta3(pE)-42 and A beta3(pE)-40 have not previously been examined. After initial solubilization and 1-week preaggregation of each peptide at 37 degrees C and pH 7.4, the toxicity of 5-50 microM A beta3(pE)-42 was similar to that of A beta1-42. Moreover, the toxicity of A beta3(pE)-40 paralleled that induced by A beta1-40 in both 1 day in vitro (DIV) cortical and 7 DIV hippocampal cells. Circular dichroism spectra did not reveal major differences in secondary structure between aged A beta1-42, A beta3(pE)-42, A beta3(pE)-40, and A beta1-40 or freshly solubilized forms of these peptides. Overall, the data indicate that the loss of the two N-terminal amino acids and the cyclization of glutamate at position 3 do not alter the extracellular toxicity of A beta.  相似文献   

13.
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   

14.
The branch specificity of Escherichia coli beta-galactosidase (EC 3.2.1.23) was studied by analyzing the cleavage of the branched hexasaccharide Gal beta 1-4GlcNAc beta 1-3(Gal beta 1-4GlcNAc beta 1-6)[14C(U)]Gal beta 1-4GlcNAc (1). This hexasaccharide was cleaved to pentasaccharides Gal beta 1-4GlcNAc beta 1-3(GlcNAc beta 1-6) [14C(U)]Gal beta 1-4GlcNAc (3) and GlcNAc beta 1-3(Gal-beta 1-4GlcNAc beta 1-6) [14C(U)]Gal beta 1-4GlcNAc (4) without any appreciable branch specificity. Even the further conversions of the pentasaccharides 3 and 4 into the tetrasaccharide GlcNAc beta 1-3(GlcNAc beta 1-6)[14C(U)]Gal beta 1-4GlcNAc seemed to proceed at similar rates, without any appreciable branch specificity. In marked contrast to the hexasaccharide 1, the pentasaccharide Gal beta 1-4GlcNAc beta 1-3(Gal beta 1-4GlcNAc beta 1-6)[14C(U)]Gal (2), missing the reducing end GlcNAc, is known to be cleaved selectively at the 6-branch; this finding was confirmed in the present study. The different behaviour of hexasaccharide 1 and pentasaccharide 2 reflects differences in the reactivity of their 6-branches; the preferred conformations of these closely related molecules may be quite different.  相似文献   

15.
H Lin  Y J Zhu  R Lal 《Biochemistry》1999,38(34):11189-11196
Amyloid beta protein (A beta P) forms senile plaques in the cerebrocortical blood vessels and brain parenchyma of patients with Alzheimer's disease (AD). The nonfamilial or sporadic AD (SAD), the most prevalent form of AD, has been correlated with an increased level of 40-residue A beta P (A beta P1-40). However, very little is known about the role of A beta P1-40 in AD pathophysiology. We have examined the activity of A beta P1-40 reconstituted in phospholipid vesicles. A combined light fluorescence and atomic force microscope (AFM) was used to image the structure of reconstituted vesicles and 45Ca2+ uptake was used as an assay for calcium permeability across the vesicular membrane. Vesicles reconstituted with fresh and globular A beta P1-40 contain a significant amount of A0 beta P and exhibit strong immunofluorescence labeling with an antibody raised against the N-terminal domain of A beta P, suggesting the incorporation of A beta P1-40 peptide in the vesicular membrane. Vesicles reconstituted with A beta P1-40 exhibited a significant level of 45Ca2+ uptake. The vesicular calcium level saturated over time, showing an important ion channel characteristic. The 45Ca2+ uptake was inhibited by (i) a monoclonal antibody raised against the N-terminal region of A beta P and (ii) Zn2+. However, a reducing agent (DTT) did not inhibit the 45Ca2+ uptake, indicating that the oxidation of A beta P or its surrounding lipid molecules is not directly involved in A beta P-mediated Ca2+ uptake. These findings provide biochemical and structural evidence that fresh and globular A beta P1-40 forms calcium-permeable channels and thus may induce cellular toxicity by regulating the calcium homeostasis in nonfamilial or sporadic Alzheimer's disease.  相似文献   

16.
Multiple factors affect the thrombin-catalyzed conversion of fibrinogen to fibrin, including: fibrinopeptide (FPA and FPB) release leading to exposure of two types of polymerization domains ("A" and "B," respectively) in the central portion of the molecule, and exposure of a noncatalytic "secondary" thrombin-binding site in fibrin. Fibrinogen containing the FPA sequence but lacking the B beta 1-42 sequence ("des-(B beta 1-42)-fibrinogen"), was compared to native fibrinogen (containing both FPA and FPB) to investigate the role played by B beta 1-42 in the polymerization of alpha-fibrin (i.e. fibrin lacking FPA), to compare reptilase and thrombin cleavage of FPA from fibrinogen, and to explore the location and function of the secondary thrombin-binding site. Electron microscopy of evolving polymer structures (mu, 0.14; pH 7.4) plus turbidity measurements, showed that early thin fibril formation as well as subsequent lateral fibril associations were impaired in des-(B beta 1-42)-alpha-fibrin, thus indicating that the B beta 1-42 sequence contributes to the A polymerization site. Reptilase-activated des-(B beta 1-42)-alpha-fibrin polymerized even more slowly than thrombin-activated des-(B beta 1-42)-alpha-fibrin, differences that disappeared when repolymerization of preformed fibrin monomers was carried out. Since existing data indicate that thrombin releases FPA in a concerted manner, resulting in relatively rapid evolution of fully functional divalent alpha-fibrin monomers, it can be inferred that delayed fibrin assembly of reptilase fibrin is due to slower formation of divalent alpha-fibrin monomers. Thrombin-activated des-(B beta 1-42)-alpha-fibrin polymerized more rapidly at low ionic strength (mu, 0.04) than did native alpha,beta-fibrin, a reversal of their behavior at physiological ionic strength (mu, 0.14). Concomitant measurement of FPA release revealed modest slowing of release at low ionic strength from des-(B beta 1-42)-fibrinogen (t1/2, 36.5 versus 21.5 min) and marked slowing from native fibrinogen (t1/2, 138 versus 22.2 min). This behavior correlated with increased thrombin binding to native alpha,beta-fibrin at low ionic strength, coupled with weak thrombin binding to des-(B beta 1-42)-alpha-fibrin, and indicates that secondary thrombin binding plays an important role in regulating thrombin diffusion and catalytic activity. Des-(B beta 1-42)-fibrinogen lacks or has a markedly defective secondary thrombin-binding site, from which we conclude that the B beta 15-42 sequence in fibrin plays a major role in forming or providing this site.  相似文献   

17.
The effect of aurovertin on the binding parameters of ADP and ATP to native F1 from beef heart mitochondria in the presence of EDTA has been explored. Three exchangeable sites per F1 were titrated by ADP and ATP in the absence or presence of aurovertin. Curvilinear Scatchard plots for the binding of both ADP and ATP were obtained in the absence of aurovertin, indicating one high affinity site (Kd for ADP = 0.6-0.8 microM; Kd for ATP = 0.3-0.5 microM) and two lower affinity sites (Kd for ADP = 8-10 microM; Kd for ATP = 7-10 microM). With a saturating concentration of aurovertin capable of filling the three beta subunits of F1, the curvilinearity of the Scatchard plots was decreased for ATP binding and abolished for ADP binding, indicating homogeneity of ADP binding sites in the F1-aurovertin complex (Kd for ADP = 2 microM). When only the high affinity aurovertin site was occupied, maximal enhancement of the fluorescence of the F1-aurovertin complex was attained with 1 mol of ADP bound per mol of F1 and maximal quenching for 1 mol of ATP bound per mol of F1. When the F1-aurovertin complex was incubated with [3H]ADP followed by [14C]ATP, full fluorescence quenching was attained when ATP had displaced the previously bound ADP. In the case of the isolated beta subunit, both ADP and ATP enhanced the fluorescence of the beta subunit-aurovertin complex. The Kd values for ADP and ATP in the presence of EDTA were 0.6 mM and 3.7 mM, respectively; MgCl2 decreased the Kd values to 0.1 mM for both ADP and ATP. It is postulated that native F1 possesses three equivalent interacting nucleotide binding sites and exists in two conformations which are in equilibrium and recognize either ATP (T conformation) or ADP (D conformation). The negative interactions between the nucleotide binding sites of F1 are strongest in the D conformation. Upon addition of aurovertin, the site-site cooperativity between the beta subunits of F1 is decreased or even abolished.  相似文献   

18.
We examine the potential antioxidant activity and the immune function of new epicatechin conjugates obtained by depolymerization of grape polymeric flavanols in the presence of cysteamine or cysteine. When incubated with an erythrocyte suspension, flavanols protected the erythrocyte membrane from hemolysis induced by 2,2'-azo-bis(2-amidinopropane)dihydrochloride (AAPH), an azo free radical initiator. The inhibitory effect was concentration-dependent and the IC50 was 119.8 microM for epicatechin, and 74.9 and 89.4 microM for the cysteine and cysteamine derivatives, respectively. These compounds were tested for their antioxidant activity and their capacity to modulate interleukin-1beta (IL-1beta), which is currently considered to be the major cytokine factor influencing the acute phase of the inflammatory response. At concentrations up to 20 microM, epicatechin and its derivatives inhibited the production of IL-1beta in whole blood incubated in the presence of E. coli lipopolysaccharide (LPS), in a concentration-dependent manner. The most active compound was the cysteamine derivative.  相似文献   

19.
Small model peptides containing N-terminal methionine are reported to form sulfur-centered-free radicals that are stabilized by the terminal N atom. To test whether a similar chemistry would apply to a disease-relevant longer peptide, Alzheimer's disease (AD)-associated amyloid beta-peptide 1-42 was employed. Methionine at residue 35 of this 42-mer has been shown to be a key amino acid residue involved in amyloid beta-peptide 1-42 [A beta1-42]-mediated toxicity and therefore, the pathogenesis of AD. Previous studies have shown that mutation of the methionine residue to norleucine abrogates the oxidative stress and neurotoxic properties of A beta(1-42). In the current study, we examined if the position of methionine at residue 35 is a criterion for toxicity. In doing so, we tested the effects of moving methionine to the N-terminus of the peptide in a synthetic peptide, A beta(1-42)D1M, in which methionine was substituted for aspartic acid at the N-terminus of the peptide and all subsequent residues from D1 to L34 were shifted one position towards the carboxy-terminus. A beta(1-42)D1M exhibited oxidative stress and neurotoxicity properties similar to those of the native peptide, A beta(1-42), all of which are inhibited by the free radical scavenger Vitamin E, suggesting that reactive oxygen species may play a role in the A beta-mediated toxicity. Additionally, substitution of methionine at the N-terminus by norleucine, A beta(1-42)D1Nle, completely abrogated the oxidative stress and neurotoxicity associated with the A beta(1-42)D1M peptide. The results of this study validate the chemistry reported for short peptides with N-terminal methionines in a disease-relevant peptide.  相似文献   

20.
Guanidine x HCl (GdnHCl)-induced unfolding of tetrameric N(5)-(L-1-carboxyethyl)-L-ornithine synthase (CEOS; 141,300 M(r)) from Lactococcus lactis at pH 7.2 and 25 degrees C occurred in several phases. The enzyme was inactivated at approximately 1 M GdnHCl. A time-, temperature-, and concentration-dependent formation of soluble protein aggregates occurred at 0.5-1.5 M GdnHCl due to an increased exposure of apolar surfaces. A transition from tetramer to unfolded monomer was observed between 2 and 3.5 M GdnHCl (without observable dimer or trimer intermediates), as evidenced by tyrosyl and tryptophanyl fluorescence changes, sulfhydryl group exposure, loss of secondary structure, size-exclusion chromatography, and sedimentation equilibrium data. GdnHCl-induced dissociation and unfolding of tetrameric CEOS was concerted, and yields of reactivated CEOS by dilution from 5 M GdnHCl were improved when unfolding took place on ice rather than at 25 degrees C. Refolding and reconstitution of the enzyme were optimal at 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