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 31 毫秒
1.
Deposition of insoluble fibrillar aggregates of β‐amyloid (Aβ) peptides in the brain is a hallmark of Alzheimer's disease. Apart from forming fibrils, these peptides also exist as soluble aggregates. Fibrillar and a variety of nonfibrillar aggregates of Aβ have also been obtained in vitro. Hexafluoroisopropanol (HFIP) has been widely used to dissolve Aβ and other amyloidogenic peptides. In this study, we show that the dissolution of Aβ40, 42, and 43 in HFIP followed by drying results in highly ordered aggregates. Although α‐helical conformation is observed, it is not stable for prolonged periods. Drying after prolonged incubation of Aβ40, 42, and 43 peptides in HFIP leads to structural transition from α‐helical to β‐conformation. The peptides form short fibrous aggregates that further assemble giving rise to highly ordered ring‐like structures. Aβ16–22, a highly amyloidogenic peptide stretch from Aβ, also formed very similar rings when dissolved in HFIP and dried. HFIP could not induce α‐helical conformation in Aβ16–22, and rings were obtained from freshly dissolved peptide. The rings formed by Aβ40, 42, 43, and Aβ16–22 are composed of the peptides in β‐conformation and cause enhancement in thioflavin T fluorescence, suggesting that the molecular architecture of these structures is amyloid‐like. Our results clearly indicate that dissolution of Aβ40, 42 and 43 and the amyloidogenic fragment Aβ16–22 in HFIP results in the formation of annular amyloid‐like structures. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
Aggregation of Aβ peptides is a seminal event in Alzheimer's disease. Detailed understanding of the Aβ assembly process would facilitate the targeting and design of fibrillogenesis inhibitors. Here, conformational studies using FTIR spectroscopy are presented. As a model peptide, the 11–28 fragment of Aβ was used. This model peptide is known to contain the core region responsible for Aβ aggregation. The structural behavior of the peptide during aggregation provoked by the addition of water to Aβ(11–28) solution in hexafluoroisopropanol was compared with the properties of its variants corresponding to natural, clinically relevant mutants at positions 21–23 (A21G, E22K, E22G, E22Q and D23N). The results showed that the aggregation of the peptides proceeds via a helical intermediate, and it is possible that the formation of α‐helical structures is preceded by creation of 310‐helix/310‐turn structures. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

3.
The crystal structure of a tripeptide Boc‐Leu‐Val‐Ac12c‐OMe ( 1 ) is determined, which incorporates a bulky 1‐aminocyclododecane‐1‐carboxylic acid (Ac12c) side chain. The peptide adopts a semi‐extended backbone conformation for Leu and Val residues, while the backbone torsion angles of the Cα,α‐dialkylated residue Ac12c are in the helical region of the Ramachandran map. The molecular packing of 1 revealed a unique supramolecular twisted parallel β‐sheet coiling into a helical architecture in crystals, with the bulky hydrophobic Ac12c side chains projecting outward the helical column. This arrangement resembles the packing of peptide helices in crystal structures. Although short oligopeptides often assemble as parallel or anti‐parallel β‐sheet in crystals, twisted or helical β‐sheet formation has been observed in a few examples of dipeptide crystal structures. Peptide 1 presents the first example of a tripeptide showing twisted β‐sheet assembly in crystals. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
Short alphahelical peptide sequences were stabilized through Glaser‐Hay couplings of propargylated l ‐ and/or d ‐serine residues at positions i and i+7. NMR analysis confirmed a full stabilization of the helical structure when a d ‐Ser (i), l ‐Ser (i+7) combination was applied. In case two l ‐Ser residues were involved in the cyclization, the helical conformation is disrupted outside the peptide's macrocycle.  相似文献   

5.
The crystal structure of the triple‐helical peptide (Pro‐Hyp‐Gly)3‐Pro‐Arg‐Gly‐(Pro‐Hyp‐Gly)4 (POG3‐PRG‐POG4) was determined at 1.45 Å resolution. POG3‐PRG‐POG4 was designed to permit investigation of the side‐chain conformation of the Arg residues in a triple‐helical structure. Because of the alternative structure of one of three Arg residues, four side‐chain conformations were observed in an asymmetric unit. Among them, three adopt a ttg?t conformation and the other adopts a tg?g?t conformation. A statistical analysis of 80 Arg residues in various triple‐helical peptides showed that, unlike those in globular proteins, they preferentially adopt a tt conformation for χ1 and χ2, as observed in POG3‐PRG‐POG4. This conformation permits van der Waals contacts between the side‐chain atoms of Arg and the main‐chain atoms of the adjacent strand in the same molecule. Unlike many other host–guest peptides, in which there is a significant difference between the helical twists in the guest and the host peptides, POG3‐PRG‐POG4 shows a marked difference between the helical twists in the N‐terminal peptide and those in the C‐terminal peptide, separated near the Arg residue. This suggested that the unique side‐chain conformation of the Arg residue affects not only the conformation of the guest peptide, but also the conformation of the peptide away from the Arg residue. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 1000–1009, 2014.  相似文献   

6.
Although unglycosylated HuEpo is fully functional, it has very short serum half‐life. However, the mechanism of in vivo clearance of human Epo (HuEpo) remains largely unknown. In this study, the relative importance of protease‐sensitive sites of recombinant HuEpo (rHuEpo) has been investigated by analysis of structural data coupled with in vivo half‐life measurements. Our results identify α3‐α4 inter‐helical loop region as a target site of lysosomal protease Cathepsin L. Consistent with previously‐reported lysosomal degradation of HuEpo, these results for the first time identify cleavage sites of rHuEpo by specific lysosomal proteases. Furthermore, in agreement with the lowered exposure of the peptide backbone around the cleavage site, remarkably substitutions of residues with bulkier amino acids result in significantly improved in vivo stability. Together, these results have implications for the mechanism of in vivo clearance of the protein in humans. Proteins 2015; 83:1813–1822. © 2015 Wiley Periodicals, Inc.  相似文献   

7.
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   

8.
Self‐assembly of PAs composed of palmitic acid and several repeated heptad peptide sequences, C15H31CO‐(IEEYTKK)n‐NH2 (n = 1–4, represented by PA1–PA4), was investigated systematically. The secondary structures of the PAs were characterized by CD. PA3 and PA4 (n = 3 and 4, respectively) showed an α‐helical structure, whereas PA1 and PA2 (n = 1 and 2, respectively) did not display an α‐helical conformations under the tested conditions. The morphology of the self‐assembled peptides in aqueous medium was studied by transmission electron microscopy. As the number of heptad repeats in the PAs increased, the nanostructure of the self‐assembled peptides changed from nanofibers to nanovesicles. Changes of the secondary structures and the self‐assembly morphologies of PA3 and PA4 in aqueous medium with various cations were also studied. The critical micelle concentrations were determined using a pyrene fluorescence probe. In conclusion, this method may be used to design new peptide nanomaterials. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
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   

10.
N‐terminal modification of peptides by unnatural amino acids significantly affects their enzymatic stability, conformational properties and biological activity. Application of N‐amidino‐amino acids, positively charged under physiological conditions, can change peptide conformation and its affinity to the corresponding receptor. In this article, we describe synthesis of short peptides, containing a new building block—N‐amidino‐pyroglutamic acid. Although direct guanidinylation of pyroglutamic acid and oxidation of N‐amidino‐proline using RuO4 did not produce positive results, N‐amidino‐Glp‐Phe‐OH was synthesized on Wang polymer by cyclization of α‐guanidinoglutaric acid residue. In the course of synthesis, it was found that literature procedure of selective Boc deprotection using TMSOTf/TEA reagent is accompanied by concomitant side reaction of triethylamine alkylation by polymer linker fragment. It should be mentioned that independently from cyclization time and coupling agent (DIC or HCTU), the lactam formation was incomplete. Separation of the cyclic product from the linear precursor was achieved by HPLC in ammonium formate buffer at pH 6. HPLC analysis showed N‐amidino‐Glp‐Phe‐OH stability at acidic and physiological pH and fast ring opening in water solution at pH 9. The suggested method of N‐amidino‐Glp residue formation can be applied in the case of short peptide chains, whereas synthesis of longer ones will require fragment condensation approach. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
A combined experimental and computational study is performed for arginine side chain stacking with the protein α‐helix. Theremostability measurements of Aristaless homeodomain, a helical protein, suggest that mutating the arginine residue R106, R137 or R141, which has the guanidino side chain stacking with the peptide plane, to alanine, destabilizes the protein. The R‐PP stacking has an energy of ~0.2‐0.4 kcal/mol. This stacking interaction mainly comes from dispersion and electrostatics, based on MP2 calculations with the energy decomposition analysis. The calculations also suggest that the stacking stabilizes 2 backbone‐backbone h‐bonds (ii‐4 and i‐3→i‐7) in a cooperative way. Desolvation and electrostatic polarization are responsible for cooperativity with the ii‐4 and i‐3→i‐7 h‐bonds, respectively. This cooperativity is supported by a protein α‐helices h‐bond survey in the pdb databank where stacking shortens the corresponding h‐bond distances.  相似文献   

12.
The interaction of the Bacillus subtilis protein Mistic with the bacterial membrane and its role in promoting the overexpression of other membrane proteins are still matters of debate. In this study, we aimed to determine whether individual helical fragments of Mistic are sufficient for its interaction with membranes in vivo and in vitro. To this end, fragments encompassing each of Mistic's helical segments and combinations of them were produced as GFP‐fusions, and their cellular localization was studied in Escherichia coli. Furthermore, peptides corresponding to the four helical fragments were synthesized by solid‐phase peptide synthesis, and their ability to acquire secondary structure in a variety of lipids and detergents was studied by circular dichroism spectroscopy. Both types of experiments demonstrate that the third helical fragment of Mistic interacts only with LDAO micelles but does not partition into lipid bilayers. Interestingly, the other three helices interact with membranes in vivo and in vitro. Nevertheless, all of these short sequences can replace full‐length Mistic as N‐terminal fusions to achieve overexpression of a human G‐protein‐coupled receptor in E. coli, although with different effects on quantity and quality of the protein produced. A bioinformatic analysis of the Mistic family expanded the number of homologs from 4 to 20, including proteins outside the genus Bacillus. This information allowed us to discover a highly conserved Shine‐Dalgarno sequence in the operon mstX‐yugO that is important for downstream translation of the potassium ion channel yugO.  相似文献   

13.
The solution structure of a peptide fragment corresponding to the 38–59 region of porcine phospholipase A2 has been investigated using CD, nmr chemical shifts, and nuclear over-hauser effects (NOEs). This isolated fragment of phospholipase forms an α-helix spanning residues 38–55, very similar to the one found in the native protein, except for residues 56–58, which were helical in the crystal but found random in solution. Addition of triflouro-ethanol (TFE) merely increased helix population but it did not redefine helix limits. To investigate how the folding information, in particular that concerning eventual helix start and stop signals, was coded in this particular amino acid sequence, the helices formed by synthetic peptides reproducing sections of this phospholipase 38–59 fragment, namely 40–59, 42–59, 38–50, and 45–57, were characterized using NOEs and helix populations quantitatively evaluated on different peptide chain segments using nmr chemical shifts in two solvents (H2O and 30% TFE/H2O). A set of nmr spectra was also recorded and assigned under denaturing conditions (6Murea) to obtain reliable values for the chemical shifts of each peptide in the random state. Based on chemical shift data, it was concluded that the helix formed by the phospholipase 38–59 fragment was not abruptly, but progressively, destabilized all along its length by successive elimination of residues at the N end, while the removal of residues at the C end affected helix stability more locally and to a lesser extent. These results are consistent with the idea that there are not single residues responsible for helix initiation or helix stability, and they also evidence an asymmetry for contributions to helix stability by residues located at the two chain ends. The restriction of molecular mobility caused by linking with a disulphide bridge at Cys 51 two identical 38–59 peptide chains did not increase helix stability. The helix formed by the covalently formed homodimer was very similar in length and population to that formed by the monomer. © 1994 John Wiley & Sons, Inc.  相似文献   

14.
The amino acid Aib predisposes a peptide to be helical with context‐dependent preference for either 310‐ or α‐ or a mixed helical conformation. Short peptides also show an inherent tendency to be unfolded. To characterize helical and unfolded states adopted by water‐soluble Aib‐containing peptides, the conformational preference of Ac‐Ala‐Aib‐Ala‐Lys‐Ala‐Aib‐Lys‐Ala‐Lys‐Ala‐Aib‐Tyr‐NH2 was determined by CD, NMR and MD simulations as a function of temperature. Temperature‐dependent CD data indicated the contribution of two major components, each an admixture of helical and extended/polyproline II structures. Both right‐ and left‐handed helical conformations were detected from deconvolution of CD data and 13C NMR experiments. The presence of a helical backbone, more pronounced at the N‐terminal, and a temperature‐induced shift in α‐helix/310‐helix equilibrium, more pronounced at the C‐terminal, emerged from NMR data. Starting from polyproline II, the N‐terminal of the peptide folded into a helical backbone in MD simulations within 5 ns at 60°C. Longer simulations showed a mixed‐helical backbone to be stable over the entire peptide at 5°C while at 60°C the mixed‐helix was either stable at the N‐terminus or occurred in short stretches through out the peptide, along with a significant population of polyproline II. Our results point towards conformational heterogeneity of water‐soluble Aib‐based peptide helices and the associated subtleties. The problem of analyzing CD and NMR data of both left‐ and right‐handed helices are discussed, especially the validity of the ellipticity ratio [θ]222/[θ]207, as a reporter of α‐/310‐ population ratio, in right‐ and left‐handed helical mixtures. Proteins 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

15.
Earlier studies have shown that the helical content of α‐helical peptide decreases upon its interaction with carbon nanotube (CNT). Further, the length of the α‐helix varies from few residues in the small globular protein to several number of residues in structural and membrane proteins. In structural and membrane proteins, helices are widely present as the supercoil i.e., helical bundles. Thus, in this study, the length‐dependent interaction pattern of α‐helical peptides with CNT and the stability of isolated α‐helical fragment versus supercoiled helical bundle upon interaction with CNT have been investigated using classical molecular dynamics (MD) simulation. Results reveal that the disruption in the helical motif on interaction with CNT is directly proportional to the length of the helix. Also it is found that the shorter helix does not undergo noticeable changes in the helicity upon adsorption with CNT. On the other hand, helicity of longer peptides is considerably affected by its interaction with CNT. In contrast to the known fact that the stability of the helix increases with its length, the disruption in the helical peptide increases with its length upon its interaction with CNT. Comparison of results shows that structural changes in the isolated helical fragment are higher than that in supercoiled helix. In fact, helical chain in supercoiled bundle does not undergo significant changes in the helicity upon interaction with CNT. Both the length of the helical peptide and the inherent stability of the helical unit in the supercoiled helix influence the interaction pattern with the CNT. © 2012 Wiley Periodicals, Inc. Biopolymers 99: 357–369, 2013.  相似文献   

16.
We have firstly demonstrated the renaturation process of dissociated single chains of lentinan (s‐LNT) and the variable conformations of the renatured LNT (r‐LNT). The results from ultrasensitive differential scanning calorimetry and circular dichroism revealed that the variable structures including perfect triple helix, defective triple helix containing duplex segment, and single chains occurred in the renaturation of s‐LNT, depending on the renaturation time, solvent composition, molecular weight, and the mode of renaturation. When water was added into s‐LNT/dimethylsulfoxide (DMSO) to reach 95% (v/v), the classic low‐temperature intra‐triple‐helical conformational transition at ~10°C (T1) appeared within 4 h, indicative of a rapid reconstruction of triple helical structure. Besides, one newly endothermic peak at ~43°C (T2) simultaneously occurred, which was first ascribed to the melting of duplex segment in the imperfect triplex. The duplex stretches disappeared when DMSO reached 50%, in which single chains coexisted with triplex. Moreover, the duplex segment disappeared by slowly dropping water into s‐LNT/DMSO. This work suggested that the structure of r‐LNT could be controllable, and provided important information for their successful development and application in polymer and life science. © 2012 Wiley Periodicals, Inc. Biopolymers 97:988–997, 2012.  相似文献   

17.
A significant fraction of the amino acids in proteins are alpha helical in conformation. Alpha helices in globular proteins are short, with an average length of about twelve residues, so that residues at the ends of helices make up an important fraction of all helical residues. In the middle of a helix, H-bonds connect the NH and CO groups of each residue to partners four residues along the chain. At the ends of a helix, the H-bond potential of the main chain remains unfulfilled, and helix capping interactions involving bonds from polar side chains to the NH or CO of the backbone have been proposed and detected. In a study of synthetic helical peptides, we have found that the sequence Ser-Glu-Asp-Glu stabilizes the alpha helix in a series of helical peptides with consensus sequences. Following the report by Harper and Rose, which identifies SerXaaXaaGlu as a member of a class of common motifs at the N termini of alpha helices in proteins that they refer to as “capping boxes,” we have reexamined the side chain–main chain interactions in a varient sequence using 1H NMR, and find that the postulated reciprocal side chain-backbone bonding between the first Ser and last Glu side chains and their peptide NH partners can be resolved: Deletion of two residues N terminal to the Ser-Glu-Asp-Glu sequence in these peptides has no effect on the initiation of helical structure, as defined by two-dimensional (2D) NMR experiments on this variant. Thus the capping box sequence Ser-Glu-Asp-Glu inhibits N terminal fraying of the N terminus of alpha helix in these peptides, and shows the side chain–main chain interactions proposed by Harper and Rose. It thus acts as a helix initiating signal. Since normal a helix cannot propagate beyond the N terminus of this structure, the box acts as a termination signal in this direction as well. © 1994 John Wiley & Sons, Inc.  相似文献   

18.
The β‐subunit of the human chorionic gonadotropin (hCG) hormone, which is believed to be related to certain types of cancer, contains three hairpin‐like fragments. To investigate the role of β‐hairpin formation in the early stages of the hCGβ folding, a 28‐residue peptide with the sequence RDVRFESIRLPGSPRGVNPVVSYAVALS, corresponding to the H3‐β hairpin fragment (residues 60–87) of the hCGβ subunit, was studied under various conditions using three optical spectroscopic methods: Fourier transform ir spectroscopy, electronic CD, and vibrational CD. Environmental conditions are critical factors for formation of secondary structure in this peptide. TFE : H2O mixed solvents induced helical formation. Formation of β‐structure in this peptide, which may be related to the native β‐hairpin formation in the intact hormone, was found to be induced only under conditions such as high concentration, high temperature, and the presence of nonmicellar sodium dodecyl sulfate concentrations. These findings support a protein folding mechanism for the hCGβ subunit in which an initial hydrophobic collapse, which increases intermolecular interactions in hCGβ, is needed to induce the H3‐β hairpin formation. © 1999 John Wiley & Sons, Inc. Biopoly 50: 413–423, 1999  相似文献   

19.
Two analogs of the ten‐amino acid residue, membrane‐active lipopeptaibiotic trichogin GA IV, mono‐labeled with 4‐cyano‐α‐methyl‐L ‐phenylalanine, a potentially useful fluorescence and IR absorption probe of the local microenvironment, were synthesized by the solid‐phase methodology and conformationally characterized. The single modification was incorporated either at the N‐terminus (position 1) or near the C‐terminus (position 8) of the peptide main chain. In both cases, the replaced amino acid was the equally helicogenic α‐aminoisobutyric acid (Aib) residue. We performed a solution conformational analysis by use of FT‐IR absorption, CD, and 2D‐NMR spectroscopies. The results indicate that both labeled analogs essentially maintain the overall helical propensity of the naturally occurring lipopeptaibiotic. Peptide? membrane interactions were assessed by fluorescence and ATR‐IR absorption techniques. Analogies and differences between the two peptides were highlighted. Taken together, our data confirm literature results that some of the spectroscopic parameters of the 4‐cyanobenzyl chromophore are sensitive markers of the local microenvironment.  相似文献   

20.
α‐melanocyte stimulating hormone (α‐MSH) is a tridecapeptide fragment of pro‐opiomelanocortin (POMC) with broad effects on appetite, skin pigmentation, hormonal regulation, and potential roles in both inflammation and autoimmunity. The use of this peptide as an anti‐inflammatory agent is limited by its low selectivity between the melanocortin receptors, susceptibility to proteolytic degradation, and rapid clearance from circulation. A retro‐inverso (RI) sequence of α‐MSH was characterized for receptor activity and resistance to protease. This peptide demonstrated surprisingly high selectivity for binding the melanocortin receptor 1 (MC1R). However, RI‐α‐MSH exhibited a diminished binding affinity for MC1R compared to α‐MSH. Mapping of the residues critical for agonist activity, receptor binding, and selectivity by alanine scanning, identified the same critical core tetrapeptide required for the native peptide. Modest improvements in affinity were obtained by conservative changes employing non‐natural amino acids and substitution of the C‐terminal sequence with a portion of a MC1R ligand peptide previously identified by phage display. Recombination of these elements yielded a peptide with an identical Ki as α‐MSH at MC1R and a lower EC50 in Mel‐624 melanoma cells. A number of other structural modifications of the RI peptide were found to differ in effect from those reported for the L ‐form α‐MSH, suggesting a significantly altered interaction with the MC1R. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  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