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 0 毫秒
1.
The 173–195 segment corresponding to the helix 2 of the globular PrP domain is a good candidate to be one of the several ‘spots’ of intrinsic structural flexibility, which might induce local destabilization and concur to protein transformation, leading to aggregation‐prone conformations. Here, we report CD and NMR studies on the α2‐helix‐derived peptide of maximal length (hPrP[180–195]) that is able to exhibit a regular structure different from the prevalently random arrangement of other α2‐helix‐derived peptides. This peptide, which has previously been shown to be affected by buffer composition via the ion charge density dependence typical of Hofmeister effects, corresponds to the C‐terminal sequence of the PrPC full‐length α2‐helix and includes the highly conserved threonine‐rich 188–195 segment. At neutral pH, its conformation is dominated by β‐type contributions, which only very strong environmental modifications are able to modify. On TFE addition, an increase of α‐helical content can be observed, but a fully helical conformation is only obtained in neat TFE. However, linking of the 173–179 segment, as occurring in wild‐type and mutant peptides corresponding to the full‐length α2‐helix, perturbs these intrinsic structural propensities in a manner that depends on whether the environment is water or TFE. Overall, these results confirm that the 180–195 parental region in hPrPC makes a strong contribution to the chameleon conformational behavior of the segment corresponding to the full‐length α2‐helix, and could play a role in determining structural rearrangements of the entire globular domain. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
3.
On consideration that intrinsic structural weakness could affect the segment spanning the α2‐helical residues 173–195 of the PrP, we have investigated the conformational stabilities of some synthetic Ala‐scanned analogs of the peptide derived from the 180–195 C‐terminal sequence, using a novel approach whose theoretical basis originates from protein thermodynamics. Even though a quantitative comparison among peptides could not be assessed to rank them according to the effect caused by single amino acid substitution, as a general trend, all peptides invariably showed an appreciable preference for an α‐type organization, consistently with the fact that the wild‐type sequence is organized as an α‐helix in the native protein. Moreover, the substitution of whatever single amino acid in the wild‐type sequence reduced the gap between the α‐ and the β‐propensity, invariably enhancing the latter, but in any case this gap was larger than that evaluated for the full‐length α2‐helix‐derived peptide. It appears that the low β‐conformation propensity of the 180–195 region depends on the simultaneous presence of all of the Ala‐scanned residues, indirectly confirming that the N‐terminal 173–179 segment could play a major role in determining the chameleon conformational behavior of the entire 173–195 region in the PrP. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
A fundamental question in protein science is what is the intrinsic propensity for an amino acid to be in an α-helix, β-sheet, or other backbone dihedral angle (-ψ) conformation. This question has been hotly debated for many years because including all protein crystal structures from the protein database, increases the probabilities for α-helical structures, while experiments on small peptides observe that β-sheet-like conformations predominate. We perform molecular dynamics (MD) simulations of a hard-sphere model for Ala dipeptide mimetics that includes steric interactions between nonbonded atoms and bond length and angle constraints with the goal of evaluating the role of steric interactions in determining protein backbone conformational preferences. We find four key results. For the hard-sphere MD simulations, we show that (1) β-sheet structures are roughly three and half times more probable than α-helical structures, (2) transitions between α-helix and β-sheet structures only occur when the backbone bond angle τ (N–Cα–C) is greater than 110°, and (3) the probability distribution of τ for Ala conformations in the “bridge” region of-ψ space is shifted to larger angles compared to other regions. In contrast, (4) the distributions obtained from Amber and CHARMM MD simulations in the bridge regions are broader and have increased τ compared to those for hard sphere simulations and from high-resolution protein crystal structures. Our results emphasize the importance of hard-sphere interactions and local stereochemical constraints that yield strong correlations between -ψ conformations and τ.  相似文献   

5.
The conformation of oligomers of β‐amino acids of the general type Ac‐[β‐Xaa]n‐NHMe (β‐Xaa = β‐Ala, β‐Aib, and β‐Abu; n = 1–4) was systematically examined at different levels of ab initio molecular orbital theory (HF/6‐31G*, HF/3‐21G). The solvent influence was considered employing two quantum‐mechanical self‐consistent reaction field models. The results show a wide variety of possibilities for the formation of characteristic elements of secondary structure in β‐peptides. Most of them can be derived from the monomer units of blocked β‐peptides with n = 1. The stability and geometries of the β‐peptide structures are considerably influenced by the side‐chain positions, by the configurations at the Cα‐ and Cβ‐atoms of the β‐amino acid constituents, and especially by environmental effects. Structure peculiarities of β‐peptides, in particular those of various helix alternatives, are discussed in relation to typical elements of secondary structure in α‐peptides. © 1999 John Wiley & Sons, Inc. Biopoly 50: 167–184, 1999  相似文献   

6.
Oxysterols, such as 7β‐hydroxy‐cholesterol (7β‐OH) and cholesterol‐5β,6β‐epoxide (β‐epoxide), may have a central role in promoting atherogenesis. This is thought to be predominantly due to their ability to induce apoptosis in cells of the vascular wall and in monocytes/macrophages. Although there has been extensive research regarding the mechanisms through which oxysterols induce apoptosis, much remains to be clarified. Given that experimental evidence has long associated alterations of calcium (Ca2+) homeostasis to apoptotic cell death, the aim of the present study was to determine the influence of intracellular Ca2+ changes on apoptosis induced by 7β‐OH and β‐epoxide. Ca2+ responses in differentiated U937 cells were assessed by epifluorescence video microscopy, using the ratiometric dye fura‐2. Over 15‐min exposure of differentiated U937 cells to 30 μM of 7β‐OH induced a slow but significant rise in fura‐2 ratio. The Ca2+ channel blocker nifedipine and the chelating agent EGTA blocked the increase in cytoplasmic Ca2+. Moreover, dihydropyridine (DHP) binding sites identified with BODIPY‐FLX‐DHP were blocked following pretreatment with nifedipine, indicating that the influx of Ca2+ occurred through L‐type channels. However, following long‐term incubation with 7β‐OH, elevated levels of cytoplasmic Ca2+ were not maintained and nifedipine did not provide protection against apoptotic cell death. Our results indicate that the increase in Ca2+ may be an initial trigger of 7β‐OH–induced apoptosis, but following chronic exposure to the oxysterol, the influence of Ca2+ on apoptotic cell death appears to be less significant. In contrast, Ca2+ did not appear to be involved in β‐epoxide–induced apoptosis. © 2009 Wiley Periodicals, Inc. J Biochem Mol Toxicol 23:324–332, 2009; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/jbt.20295  相似文献   

7.
The biological underpinnings linking stress to Alzheimer's disease (AD) risk are poorly understood. We investigated how corticotrophin releasing factor (CRF), a critical stress response mediator, influences amyloid‐β (Aβ) production. In cells, CRF treatment increases Aβ production and triggers CRF receptor 1 (CRFR1) and γ‐secretase internalization. Co‐immunoprecipitation studies establish that γ‐secretase associates with CRFR1; this is mediated by β‐arrestin binding motifs. Additionally, CRFR1 and γ‐secretase co‐localize in lipid raft fractions, with increased γ‐secretase accumulation upon CRF treatment. CRF treatment also increases γ‐secretase activity in vitro, revealing a second, receptor‐independent mechanism of action. CRF is the first endogenous neuropeptide that can be shown to directly modulate γ‐secretase activity. Unexpectedly, CRFR1 antagonists also increased Aβ. These data collectively link CRF to increased Aβ through γ‐secretase and provide mechanistic insight into how stress may increase AD risk. They also suggest that direct targeting of CRF might be necessary to effectively modulate this pathway for therapeutic benefit in AD, as CRFR1 antagonists increase Aβ and in some cases preferentially increase Aβ42 via complex effects on γ‐secretase.  相似文献   

8.
β‐dystroglycan (β‐DG) is a widely expressed transmembrane protein that plays important roles in connecting the extracellular matrix to the cytoskeleton, and thereby contributing to plasma membrane integrity and signal transduction. We previously observed nuclear localization of β‐DG in cultured cell lines, implying the existence of a nuclear targeting mechanism that directs it to the nucleus instead of the plasma membrane. In this study, we delineate the nuclear import pathway of β‐DG, characterizing a functional nuclear localization signal (NLS) in the β‐DG cytoplasmic domain, within amino acids 776–782. The NLS either alone or in the context of the whole β‐DG protein was able to target the heterologous GFP protein to the nucleus, with site‐directed mutagenesis indicating that amino acids R779 and K780 are critical for NLS functionality. The nuclear transport molecules Importin (Imp)α and Impβ bound with high affinity to the NLS of β‐DG and were found to be essential for NLS‐dependent nuclear import in an in vitro reconstituted nuclear transport assay; cotransfection experiments confirmed the dependence on Ran for nuclear accumulation. Intriguingly, experiments suggested that tyrosine phosphorylation of β‐DG may result in cytoplasmic retention, with Y892 playing a key role. β‐DG thus follows a conventional Impα/β‐dependent nuclear import pathway, with important implications for its potential function in the nucleus. J. Cell. Biochem. 110: 706–717, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

9.
Determining the cause of human calcitonin (hCT) aggregation could be of help in the effort to utilize hCT for treatment of hypercalcemia. Here we report that a dimer model of hCT13‐32 aggregated to a greater degree than native hCT under aqueous 2,2,2‐trifluoroethanol conditions. Analyses using circular dichroism spectroscopy, thioflavine‐T binding assays and atomic force microscopy suggest that the α‐helical portion of hCT is important for initiation of the aggregation process, which yields long fibrils. Dimerization, which stabilizes the β‐sheet structure of hCT, enhances aggregation potency. Dimerization of hCT stabilizes the α‐helix under aqueous TFE conditions, leading to the long fibril formation. Up to now, there have been no reports of using a dimer model to investigate the properties of hCT aggregation. Our findings could potentially serve as the basis for development of novel hCT derivatives that could be utilized for treatment of hypercalcemia, as well as for development of novel therapeutics for other ailments caused by amyloid peptides. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
A single chiral cyclic α,α‐disubstituted amino acid, (3S,4S)‐1‐amino‐(3,4‐dimethoxy)cyclopentanecarboxylic acid [(S,S)‐Ac5cdOM], was placed at the N‐terminal or C‐terminal positions of achiral α‐aminoisobutyric acid (Aib) peptide segments. The IR and 1H NMR spectra indicated that the dominant conformations of two peptides Cbz‐[(S,S)‐Ac5cdOM]‐(Aib)4‐OEt ( 1) and Cbz‐(Aib)4‐[(S,S)‐Ac5cdOM]‐OMe (2) in solution were helical structures. X‐ray crystallographic analysis of 1 and 2 revealed that a left‐handed (M) 310‐helical structure was present in 1 and that a right‐handed (P) 310‐helical structure was present in 2 in their crystalline states. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
The increasing interest in click chemistry and its use to stabilize turn structures led us to compare the propensity for β‐turn stabilization of different analogs designed as mimics of the β‐turn structure found in tendamistat. The β‐turn conformation of linear β‐amino acid‐containing peptides and triazole‐cyclized analogs were compared to ‘conventional’ lactam‐ and disulfide‐bridged hexapeptide analogs. Their 3D structures and their propensity to fold in β‐turns in solution, and for those not structured in solution in the presence of α‐amylase, were analyzed by NMR spectroscopy and by restrained molecular dynamics with energy minimization. The linear tetrapeptide Ac‐Ser‐Trp‐Arg‐Tyr‐NH2 and both the amide bond‐cyclized, c[Pro‐Ser‐Trp‐Arg‐Tyr‐D ‐Ala] and the disulfide‐bridged, Ac‐c[Cys‐Ser‐Trp‐Arg‐Tyr‐Cys]‐NH2 hexapeptides adopt dominantly in solution a β‐turn conformation closely related to the one observed in tendamistat. On the contrary, the β‐amino acid‐containing peptides such as Ac‐(R)‐β3‐hSer‐(S)‐Trp‐(S)‐β3‐hArg‐(S)‐β3‐hTyr‐NH2, and the triazole cyclic peptide, c[Lys‐Ser‐Trp‐Arg‐Tyr‐βtA]‐NH2, both specifically designed to mimic this β‐turn, do not adopt stable structures in solution and do not show any characteristics of β‐turn conformation. However, these unstructured peptides specifically interact in the active site of α‐amylase, as shown by TrNOESY and saturation transfer difference NMR experiments performed in the presence of the enzyme, and are displaced by acarbose, a specific α‐amylase inhibitor. Thus, in contrast to amide‐cyclized or disulfide‐bridged hexapeptides, β‐amino acid‐containing peptides and click‐cyclized peptides may not be regarded as β‐turn stabilizers, but can be considered as potential β‐turn inducers. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
β‐Amino acids containing hybrid peptides and β‐peptides show great potential as peptidomimetics. In this paper we describe the synthesis and affinity toward the µ‐ and δ‐opioid receptors of β‐peptides, analogues of Leu‐enkephalin, deltorphin I, dermorphin and α,β‐hybrides, analogues of deltorphin I. Substitution of α‐amino acid residues with β3homo‐amino acid residues, in general resulted in decrease of affinity to opioid receptors. However, the incorporation β3h‐D ‐Ala in position 2 or β3hPhe in position 3 of deltorphin I resulted in potent and selective ligand for δ‐opioid receptor. The NMR studies of β‐deltorphin I analogue suggest that conformational motions in the central part of the peptide backbone are partially restricted and some conformational preferences can be expected. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
14.
Two 8‐µs all‐atom molecular dynamics simulations have been performed on the two highly homologous G protein‐coupled receptor (GPCR) subtypes, β1‐ and β2‐adrenergic receptors, which were embedded in a lipid bilayer with randomly dispersed cholesterol molecules. During the simulations, cholesterol molecules accumulate to different surface regions of the two receptors, suggesting the subtype specificity of cholesterol–β‐adrenergic receptor interaction and providing some clues to the physiological difference of the two subtypes. Meanwhile, comparison between the two receptors in interacting with cholesterols shed some new light on general determinants of cholesterol binding to GPCRs. Our results indicate that although the concave surface, charged residues and aromatic residues are important, neither of these stabilizing factors is indispensable for a cholesterol interaction site. Different combinations of these factors lead to the diversified binding modes of cholesterol binding to the receptors. Our long‐time simulations, for the first time, revealed the pathway of a cholesterol molecule entering the consensus cholesterol motif (CCM) site, and the binding process of cholesterol to CCM is accompanied by a side chain flipping of the conserved Trp4.50. Moreover, the simulation results suggest that the I‐/V‐/L‐rich region on the extracellular parts of helix 6 might be an alternatively conserved cholesterol‐binding site for the class‐A GPCRs. Proteins 2014; 82:760–770. © 2013 Wiley Periodicals, Inc.  相似文献   

15.
We show that three designed cyclic d ,l ‐peptides are β‐helical in TFE—a solvent in which the archetypal β‐helical peptide, gA, is unstructured. This result represents an advance in the field of β‐helical peptide foldamers and a step toward achieving β‐helical structure under a broad range of solvent conditions. We synthesized two of the three peptides examined using an improved variant of our original CBC strategy. Here, we began with a commercially available PEG–PS composite resin prefunctionalized with the alkanesulfonamide ‘SCL’ linker and preloaded with glycine. Our new conditions avoided C‐terminal epimerization during the CBC step and simplified purification. In addition, we present results to define the scope and limitations of our CBC strategy. These methods and observations will prove useful in designing additional cyclic β‐helical peptides for applications ranging from transmembrane ion channels to ligands for macromolecular targets. Published 2014. This article is a U.S. Government work and is in the public domain in the USA.  相似文献   

16.
17.
The relationship between the conformation and biological activity of the peptide allosteric modulator of the interleukin‐1 receptor 101.10 (D ‐Arg‐D ‐Tyr‐D ‐Thr‐D ‐Val‐D ‐Glu‐D ‐Leu‐D ‐Ala‐NH2) has been studied using (R)‐ and (S)‐Bgl residues. Twelve Bgl peptides were synthesized using (R)‐ and (S)‐cyclic sulfamidate reagents derived from L ‐ and D ‐aspartic acid in an optimized Fmoc‐compatible protocol for efficient lactam installment onto the supported peptide resin. Examination of these (R)‐ and (S)‐Bgl 101.10 analogs for their potential to inhibit IL‐1β‐induced thymocyte cell proliferation using a novel fluorescence assay revealed that certain analogs exhibited retained and improved potency relative to the parent peptide 101.10. In light of previous reports that Bgl residues may stabilize type II′β‐turn‐like conformations in peptides, CD spectroscopy was performed on selected compounds to identify secondary structure necessary for peptide biological activity. Results indicate that the presence of a fold about the central residues of the parent peptide may be important for activity. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
Modulation of protein–protein interactions involved in the immune system by using small molecular mimics of the contact interfaces may lead to the blockage of the autoimmune response and the development of drugs for immunotherapy. The nonpolymorphic β‐regions, exposed to the microenvironment, of the modeled HLA‐DQ7, which is genetically linked to autoimmune diseases, were determined. Peptides 132–141 and 58–67, located at the β1 and β2 domains of HLA‐DQ7, respectively, were tested for their involvement in the interactions with CD4+ T lymphocytes. Linear, cyclic, and dimeric analogs that mimic the exposed surfaces of HLA‐DQ7 were designed and synthesized. Their immunosuppressory activities, found in the secondary, humoral immune response to sheep erythrocytes (SRBC) in mice in vitro, ranged from 11% to 53%. The significance of the total charge of the peptides, the pattern of the hydrogen bonding, and the presence of secondary structure were investigated in relation to the immunomodulatory effect of the peptides. Two dimeric analogs of the HLA‐DQ7 58–67 fragment, consisting of the two monomers covalently linked by a polyethylene glycol (PEG) spacer, able to mimic the superdimers, were also synthesized and studied. As the 58–67 segment is located at the β1 region of HLA‐DQ7, close to the major histocompatibility complex (MHC) groove, one may assume that the 58–67 peptide could accommodate the association between T‐cell receptor (TCR) and human leukocyte antigen (HLA) by activating a co‐stimulatory molecule of the TCR/HLA interaction. This hypothesis is supported by the confocal laser image of the fluorescein‐labeled 58–67 peptide and by the fact that it is an immunostimulator at low concentration. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
The introduction of non‐natural modules could provide unprecedented control over folding/unfolding behavior, conformational stability, and biological function of proteins. Success requires the interrogation of candidate modules in natural contexts. Here, expressed protein ligation is used to replace a reverse turn in bovine pancreatic ribonuclease (RNase A) with a synthetic β‐dipeptide: β2‐homoalanine–β3‐homoalanine. This segment is known to adopt an unnatural reverse‐turn conformation that contains a 10‐membered ring hydrogen bond, but one with a donor–acceptor pattern opposite to that in the 10‐membered rings of natural reverse turns. The RNase A variant has intact enzymatic activity, but unfolds more quickly and has diminished conformational stability relative to native RNase A. These data indicate that hydrogen‐bonding pattern merits careful consideration in the selection of beneficial reverse‐turn surrogates.  相似文献   

20.
Cyclic tetra‐β‐peptides (CP4s) and a bis‐CP4 were synthesized to prepare peptide nanotubes (PNTs) by molecular stacking of cyclic peptides. The addition of bis‐CP4 to the PNT preparation afforded PNT bundles increasing the direct and converse piezoelectiric coefficients, which is ascribable to bis‐CP4 stapling PNTs into the parallel alignment of PNT dipoles.  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