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 842 毫秒
1.
Incorporation of ferrocenyl group to peptides is an efficient method to alter their hydrophobicity. Ferrocenyl group can also act as an electrochemical probe when incorporated onto functional peptides. Most often, ferrocene is incorporated onto peptides post‐synthesis via amide, ester or triazole linkages. Stable amino acids containing ferrocene as a C‐linked side chain are potentially useful building units for the synthesis of ferrocene‐containing peptides. We report here an efficient route to synthesize ferrocene‐containing amino acids that are stable and can be used in peptide synthesis. Coupling of 2‐ferrocenyl‐1,3‐dithiane and iodides derived from aspartic acid or glutamic acid using n‐butyllithium leads to the incorporation of a ferrocenyl unit to the δ‐position or ε‐position of an α‐amino acid. The reduction or hydrolysis of the dithiane group yields an alkyl or an oxo derivative. The usability of the synthesized amino acids is demonstrated by incorporating one of the amino acids in both C‐terminus and N‐terminus of tripeptides in solution phase. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
In keeping with recent efforts to generate compounds for antibiotic and microbicide development, we focused on the creation of non‐natural organo‐peptide hybrids of antimicrobial peptide amides (KLK(L)nKLK‐NH2) derived from sapecin B and a self‐assembling oligoglycine organo‐peptide bolaphile containing an ω‐amino fatty acid residue. The hybrid organo‐peptide bolaphiles with two cationic KLK tripeptide motifs linked with an ω‐amino acid residue (penta‐, octa‐ or undecamethylene chain) maintained the self‐assembling properties of the root oligoglycine bolaphile. Electron microscopy clearly revealed complex supramolecular architectures for both sapecin B‐derived peptides and the hybrid analogues. FT‐IR spectroscopy indicated that the supramolecular structures were composed primarily of β‐sheets. CD revealed that the hybrid bolaphiles did not share the same secondary structures as the sapecin B peptides in solution. However, although secondary structures of antimicrobial peptides are central in the activity, the organo‐peptide bolaphiles also retained the potent antimicrobial activity of the leader sapecin B‐derived peptide against both Gram‐positive and Gram‐negative bacteria. In general, the hybrids were more selective than the sapecin B peptides, as they displayed little or no appreciable haemolytic activity. The results obtained herald a new approach for the design of purpose‐built hybrid organo‐peptide bolaphiles. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

3.
Summary. A series of insect oostatic peptides containing 3,4-dehydroproline in the C-terminal part or inside of the peptide chain was synthesized and tritiated by addition of 3H2 to double bond of 3,4-dehydroproline residue. 3H-label was introduced also into tyrosine residue of oostatic tetra- and pentapeptides by isotopic exchange of benzyl β-hydrogens. In this way, three types of tritiated peptides were prepared, different in the radiolabeled amino acid position: [3H] Tyr-Asp-Pro-Ala-OH, H-Tyr-Asp-[3H] Pro-Ala-OH, [3H] Tyr-Asp-Pro-Ala-Pro-OH, H-Tyr-Asp-[3H] Pro-Ala-Pro-OH, H-Tyr-Asp-Pro-Ala-[3H] Pro-OH, H-Tyr-Asp-Pro-Ala-Pro5-[3H] Pro-OH and H-Asp-[3H] Pro-OH. These peptides made possible a highly sensitive comparative study on radioactivity incorporation into head and ovaries of the flesh fly Neobellieria bullata, which revealed this process to proceed differently. The reasons of the found differences are discussed.  相似文献   

4.
A high proportion of peptide transmitters and peptide hormones terminate their peptide chain in a C-terminal amide group which is essential for their biological activity. The specificity of an enzyme that catalyses the formation of the amide was investigated with the aid of synthetic peptide substrates. With peptides containing l-amino acids the enzyme exhibited an essential requirement for glycine in the C-terminal position; amidation did not take place with peptides that had leucine, alanine, glutamic acid, lysine or N-methylglycine at the C-terminus and a peptide extended by the attachment of lysine to the C-terminal glycine did not act as a substrate. Amidation did occur with a peptide containing C-terminal D-alanine but no reaction was detected with peptides having C-terminal, D-serine or D-leucine. In tripeptides with a neutral amino acid in the penultimate position, amidation, took place readily but the reaction was slower when this position was occupied by an acidic or a basic residue. A series of overlapping peptides with C-terminal glycine, based on partial sequences of calcitonin, underwent amidation at similar rates, indicating that the amidating enzyme recognizes only a limited sequence at the C-terminus of its substrates. The results provide evidence that the amidating enzyme has a highly compact substrate binding site.  相似文献   

5.
A relaxin‐like gonad‐stimulating peptide (RGP) in starfish was the first identified invertebrate gonadotropin responsible for final gamete maturation. An RGP ortholog was newly identified from Astropecten scoparius of the order Paxillosida. The A. scoparius RGP (AscRGP) precursor is encoded by a 354 base pair open reading frame and is a 118 amino acid (aa) protein consisting of a signal peptide (26 aa), B‐chain (21 aa), C‐peptide (47 aa), and A‐chain (24 aa). There are three putative processing sites (Lys‐Arg) between the B‐chain and C‐peptide, between the C‐peptide and A‐chain, and within the C‐peptide. This structural organization revealed that the mature AscRGP is composed of A‐ and B‐chains with two interchain disulfide bonds and one intrachain disulfide bond. The C‐terminal residues of the B‐chain are Gln‐Gly‐Arg, which is a potential substrate for formation of an amidated C‐terminal Gln residue. Non‐amidated (AscRGP‐GR) and amidated (AscRGP‐NH2) peptides were chemically synthesized and their effect on gamete shedding activity was examined using A. scoparius ovaries. Both AscRGP‐GR and AscRGP‐NH2 induced oocyte maturation and ovulation in similar dose‐dependent manners. This is the first report on a C‐terminally amidated functional RGP. Collectively, these results suggest that AscRGP‐GR and AscRGP‐NH2 act as a natural gonadotropic hormone in A. scoparius.  相似文献   

6.
The β-sheet breaker (BSB) peptides interfere with amyloid fibril assembly and used as therapeutic agents in the treatment of Alzheimer's disease (AD). In this regard, a simple yet effective in silico screening methodology was applied in the present study to evaluate a potential 867 pentapeptide library based on known BSB peptide, LPFFD, for destabilizing Aβ42 protofibrils. The molecular docking based virtual screening was used to filter out pentapeptides having binding affinities stronger than LPFFD. In the next step, binding free energies of the top 10 pentapeptides were evaluated using the MM-PBSA method. The residue-wise binding free energy analysis reveals that two pentapeptides, PVFFE, and PPFYE, bind to the surface of Aβ42 protofibril and another pentapeptide, PPFFE, bind in the core region of Aβ42 protofibril. By employing molecular dynamics simulation as a post filter for the top-hit peptides from MM-PBSA, the pentapeptides, PPFFE, PVFFE, and PPFYE, have been identified as potential BSB peptides for destabilizing Aβ42 protofibril structure. The conformational microstate analysis, a significant decrease in the β-sheet content of Aβ42 protofibril, a loss in the total number of hydrogen bonds in Aβ42 protofibril, Asp23-Lys28 salt bridge destabilization and analysis of the free energy surfaces highlight Aβ42 protofibril structure destabilization in presence of pentapeptides. Among three top-hit pentapeptides, PPFFE displayed the most potent Aβ42 protofibril destabilization effect that shifted the energy minima toward lowest value of β-sheet content as well as lowest number of hydrogen bonds in Aβ42 protofibril. The in silico screening workflow presented in the study highlight an alternative tool for designing novel peptides with enhanced BSB ability as potential therapeutic agents for AD.  相似文献   

7.
The purpose of this study was to examine the effects of structural parameters of peptides on their oxidation by DMSO, including location of cysteine, effect of adjunct group participation, molecular hydrophobicity, steric hindrance or the accessibility of thiol group and peptide conformation, on oxidation rates, dimer formation and associated side products. We designed and synthesized two series of linear cysteine‐containing analogues of human β‐defensin 3 (the C1‐peptides with cysteine at the N‐terminus residue 1, the C29‐peptides with cysteine located at residue 29 in the centre of peptide), which were used for preparation of disulphide‐linked homodimers. HPLC–ESI–MS was used to monitor the oxidation process and to characterize the molecular weights of dimers and side products of high oxidation. The formations of dimers and side products were dependent on the position of cysteines. Hydrophobicity generally rendered the thiol groups less accessible and hence exposed them to slow oxidation to form dimers (or even fail to form dimers during the timescale of observation). Molecular dynamics simulations showed that the exposure of cysteines (and sulphurs) of the C1‐peptides was much larger than for the C29‐peptides. The larger hydrophobic side chains tended to enable clustering of the side chains that sequester cysteine, particularly in the C29‐peptides, which provided a molecular explanation for the observed trends in oxidation rates. Together with molecular modelling, we propose a reaction mechanism to elucidate the oxidation results of these peptides. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
The ON acyl transfer reaction has gained significant popularity in peptide and medicinal chemistry. This reaction has been successfully applied to the synthesis of difficult sequence‐containing peptides, cyclic peptides, epimerization‐free fragment coupling and more recently, to switchable peptide polymers. Herein, we describe a related strategy to facilitate the synthesis and purification of a hydrophobic stapled peptide. The staple consists of a serine linked through an amide bond formed from its carboxylic acid function and the side chain amino group of diaminopropionic acid and through an ester bond formed from its amino group and the side chain carboxylic acid function of aspartic acid. The α‐amino group of serine was protonated during purification. Interestingly, when the peptide was placed at physiological pH, the free amino group initiated the O‐N shift reducing the staple length by one atom, leading to a more hydrophobic stapled peptide. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
Fourier transform ir spectra have been recorded for three 310‐helical and one α‐helical pentapeptides containing dehydrophenylalanine, in a thin solid film, in order to find marker bands for various secondary structures encountered in peptides containing dehydroaminoacids. The peptide solutions were deposited and dried as thin film on zinc selenide crystal surface. This convenient sampling method has provided reliable estimates of peptide secondary structure in solid state. Detailed vibrational assignments in the spectral region between 1200–1700 cm−1 are reported. In this region, peptide amide I, II, and III vibrations occur. Spectra–structure correlation has been presented based on the amide modes. Comparison of the ir spectra with available crystal structure data provides qualitative support for assignments of ir bands to 310‐helical structure and α‐helical structure in dehydrophenylalanine containing pentapeptides. Band frequency assignments for 310‐helical conformation are consistent for all three peptides. All the assignments agree closely with the theoretical predictions. The spectral differences between 310‐helical peptides and the α‐helical peptide have been highlighted. These findings demonstrate that a method based on ir spectroscopy can be developed for a useful approximation of three‐dimensional structure of dehydropeptides in solid state. © 1999 John Wiley & Sons, Inc. Biopoly 50: 595–601, 1999  相似文献   

10.
Understanding the secondary structure of peptides is important in protein folding, enzyme function, and peptide‐based drug design. Previous studies of synthetic Ala‐based peptides (>12 a.a.) have demonstrated the role for charged side chain interactions involving Glu/Lys or Glu/Arg spaced three (i, i + 3) or four (i, i + 4) residues apart. The secondary structure of short peptides (<9 a.a.), however, has not been investigated. In this study, the effect of repetitive Glu/Lys or Glu/Arg side chain interactions, giving rise to E‐R/K helices, on the helicity of short peptides was examined using circular dichroism. Short E‐R/K–based peptides show significant helix content. Peptides containing one or more E‐R interactions display greater helicity than those with similar E‐K interactions. Significant helicity is achieved in Arg‐based E‐R/K peptides eight, six, and five amino acids long. In these short peptides, each additional i + 3 and i + 4 salt bridge has substantial contribution to fractional helix content. The E‐R/K peptides exhibit a strongly linear melt curve indicative of noncooperative folding. The significant helicity of these short peptides with predictable dependence on number, position, and type of side chain interactions makes them an important consideration in peptide design.  相似文献   

11.
Fourteen tryptic peptides and nine intermediates were identified as products of trypsin digestion of reduced and S-3-(trimethylated amino) propylated lysozyme. Kinetics of the appearance and disappearance of these products were observed by monitoring the peak areas on the chromatogram. In spite of the complicated reaction pathways, kinetics of the digestion of proteins and several intermediate products show simple decay curves with a single rate constant. In this paper, the trypsin susceptibility of the individual cleavage site is defined as a hydrolytic rate constant of the susceptible peptide bond in the presence of 10 nM trypsin. The cleavage sites of unfolded lysozyme are classified into two groups in terms of the trypsin susceptibility: one has a high susceptibility (10–20 h?1) and the other a low susceptibility (1.0–2.0 h?1). In the unfolded state of lysozyme, in conclusion, the region from residues 15 to 61 has a strong resistance to trypsin digestion; on the other hand, the C-terminal half of the polypeptide chain is flexible enough to fit into the active site of trypsin. In addition, six kinds of pentapeptides were synthesized as analogues of lysozyme fragments including Arg 14, Arg 21, Lys 33, Arg 45, Arg 61, and Arg 73. Kinetics of typtic digestion of them were observed. Both kcat and KM were determined for these synthetic pentapeptides. The susceptibility of each cleavage site in pentapeptides is determined and compared with that corresponding in proteins. The susceptibility is usually higher when the susceptible peptide chain is flexible. However, susceptibilities of a few sites in proteins are lower than those in pentapeptides. This means that the peptapeptides, this means that the peptide chains tend to fold locally to prevent trypsin from binding to the sites. It was found that the sites of Arg 21 and Arg 45 are indeed resistant to trypsin, but the site of Lys 33 is not so much, although the hydrolytic rate at Lys 33 itself is extremely slow. © 1994 John Wiley & Sons, Inc.  相似文献   

12.
The bacterial cell wall consists of peptidoglycan (PG), a sturdy mesh of glycan strands cross‐linked by short peptides. This rigid structure constrains cell shape and size, yet is sufficiently dynamic to accommodate insertion of newly synthesized PG, which was long hypothesized, and recently demonstrated, to require cleavage of the covalent peptide cross‐links that couple previously inserted material. Here, we identify several genes in Vibrio cholerae that collectively are required for growth – particularly elongation – of this pathogen. V. cholerae encodes three putative periplasmic proteins, here denoted ShyA, ShyB, and ShyC, that contain both PG binding and M23 family peptidase domains. While none is essential individually, the absence of both ShyA and ShyC results in synthetic lethality, while the absence of ShyA and ShyB causes a significant growth deficiency. ShyA is a D,d ‐endopeptidase able to cleave most peptide chain cross‐links in V. cholerae's PG. PG from a ?shyA mutant has decreased average chain length, suggesting that ShyA may promote removal of short PG strands. Unexpectedly, ShyA has little activity against muropeptides containing pentapeptides, which typically characterize newly synthesized material. ShyA's substrate‐dependent activity may contribute to selection of cleavage sites in PG, whose implications for the process of side‐wall growth are discussed.  相似文献   

13.
 Previous studies have defined two different peptide binding motifs specific for HLA-A * 0101. These motifs are characterized by the presence of tyrosine (Y) at the C-termini of 9-mer and 10-mer peptides, and either a small polar or hydrophobic (S, T, M) residue in position 2, or a negatively charged (D or E) residue in position 3. In this study, the structural requirements for peptide binding to A * 0101 have been further analyzed by examining the binding capacity of large sets of peptides corresponding to naturally occurring sequences which bore one or the other of these two A * 0101-specific motifs. By correlating the presence of specific residue types at each position along the peptide sequence with increased (or decreased) binding affinity, the prominent influence of secondary anchor residues was revealed. In most cases, the two anchors in positions 2 and 3 appear to act synergistically. With the exception of the DE3 submotif in 9-mer peptides, a positive role for aromatic residues in position 1 and the center of the peptide (positions 4 or 5 of 9- or 10-mer peptides, respectively), and proline at C-3, were also consistently detected. However, secondary anchor residues also appear to differ significantly between the two different submotifs, demonstrating that A * 0101 can utilize alternative modes in binding its peptide ligands. According to these analyses, specific refined submotifs were also established, and their merit verified by independent sets of potential A * 0101 binding peptides. Besides providing useful insight into the nature of the interaction of the A * 0101 allele with its peptide ligands, such refined motifs should also facilitate accurate prediction of potential A * 0101-restricted peptide epitopes. Received: 16 July 1996 / Revised: 18 September 1996  相似文献   

14.
An isolated uncharged hydrogen bond acceptor such as the carbonyl functionality of an aldehyde or a keto group is absent in natural amino acids. Although glutamine and asparagine are known to hydrogen bond through the amide carbonyl group in their side chains, they also possess the amide ? NH2 group, which can act as a hydrogen bond donor. This makes the structural study of peptides containing an oxo residue, with an isolated carbonyl group in the side chain, interesting. Here, we report the synthesis of δ‐ and ε‐oxo amino acids and their incorporation into oligopeptides as the N‐terminal residue. The resultant oxo peptides were extensively studied using X‐ray crystallography to understand the interactions offered by the oxo group in peptide crystals. We find that the oxo groups are capable of providing additional hydrogen bonding opportunities to the peptides, resulting in increased intermolecular interactions in crystals. The study thus offers avenues for the utilization of oxo residues to introduce intermolecular interactions in synthetic peptides.  相似文献   

15.
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   

16.
Inactivation of sodium channels is thought to be mediated by an inactivation gate formed by the intracellular loop connecting domains III and IV. A hydrophobic motif containing the amino acid sequence isoleucine, phenylalanine, and methionine (IFM) is required for the inactivation process. Peptides containing the IFM motif, when applied to the cytoplasmic side of these channels, produce two types of block: fast block, which resembles the inactivation process, and slow, use-dependent block stimulated by strong depolarizing pulses. Fast block by the peptide ac-KIFMK-NH2, measured on sodium channels whose inactivation was slowed by the α-scorpion toxin from Leiurus quinquestriatus (LqTx), was reversed with a time constant of 0.9 ms upon repolarization. In contrast, control and LqTx-modified sodium channels were slower to recover from use-dependent block. For fast block, linear peptides of three to six amino acid residues containing the IFM motif and two positive charges were more effective than peptides with one positive charge, whereas uncharged IFM peptides were ineffective. Substitution of the IFM residues in the peptide ac-KIFMK-NH2 with smaller, less hydrophobic residues prevented fast block. The positively charged tripeptide IFM-NH2 did not cause appreciable fast block, but the divalent cation IFM-NH(CH2)2NH2 was as effective as the pentapeptide ac-KIFMK-NH2. The constrained peptide cyclic KIFMK containing two positive charges did not cause fast block. These results indicate that the position of the positive charges is unimportant, but flexibility or conformation of the IFM-containing peptide is important to allow fast block. Slow, use-dependent block was observed with IFM-containing peptides of three to six residues having one or two positive charges, but not with dipeptides or phenylalanine-amide. In contrast to its lack of fast block, cyclic KIFMK was an effective use-dependent blocker. Substitutions of amino acid residues in the tripeptide IFM-NH2 showed that large hydrophobic residues are preferred in all three positions for slow, use-dependent block. However, substitution of the large hydrophobic residue diphenylalanine or the constrained residues phenylglycine or tetrahydroisoquinoline for phe decreased potency, suggesting that this phe residue must be able to enter a restricted hydrophobic pocket during the binding of IFM peptides. Together, the results on fast block and slow, use-dependent block indicate that IFM peptides form two distinct complexes of different stability and structural specificity with receptor site(s) on the sodium channel. It is proposed that fast block represents binding of these peptides to the inactivation gate receptor, while slow, use-dependent block represents deeper binding of the IFM peptides in the pore.  相似文献   

17.
We have recently reported a series of synthetic anticancer heptapeptides (H‐KKWβ2,2WKK‐NH2) containing a central achiral and lipophilic β2,2‐amino acid that display low toxicity against non‐malignant cells and high proteolytic stability. In the present study, we have further investigated the effects of increasing the rigidity and amphipathicity of two of our lead heptapeptides by preparing a series of seven to five residue cyclic peptides containing the two most promising β2,2‐amino acid derivatives as part of the central lipophilic core. The peptides were tested for anticancer activity against human Burkitt's lymphoma (Ramos cells), haemolytic activity against human red blood cells (RBC) and cytotoxicity against healthy human lung fibroblast cells (MRC‐5). The results demonstrated a considerable increase in anticancer potency following head‐to‐tail peptide cyclization, especially for the shortest derivatives lacking a tryptophan residue. High‐resolution NMR studies and molecular dynamics simulations together with an annexin‐V‐FITC and propidium iodide fluorescent assay showed that the peptides had a membrane disruptive mode of action and that the more potent peptides penetrated deeper into the lipid bilayer. The need for new anticancer drugs with novel modes of action is demanding, and development of short cyclic anticancer peptides with an overall rigidified and amphipathic structure is a promising approach to new anticancer agents. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
A set of analogues of the 14‐residue peptaibol tylopeptin B, containing the stable free‐radical 4‐amino‐1‐oxyl‐2,2,6,6,‐tetramethylpiperidine‐4‐carboxylic acid (TOAC) at one or two selected positions, was synthesized by the solid‐phase methodology. A solution conformational analysis performed by FTIR absorption and CD suggests that, in membrane‐mimicking solvents, the labeled tylopeptin B analogues preserve the helical propensity of the parent peptide, with a preference for the α‐helix or the 310‐helix type depending upon the nature of the solvent. In aqueous environment, the spin‐labeled analogues present a higher content of helical conformation as a consequence of the strong helix promoter effect of the conformationally constrained TOAC residue. We observed a progressive increase of the quenching effect of the nitroxyl radical on the fluorescence of the N‐terminal tryptophan as TOAC replaces the Aib residue at positions 13, 8, and 4, respectively. A membrane permeabilization assay performed on two selected analogues, TOAC8‐ and TOAC13‐tylopeptin B, showed that the labeled peptides exhibit membrane‐modifying properties comparable with those of the natural peptaibiotic. We conclude that our TOAC paramagnetic analogues of tylopeptin B are good models for a detailed ESR investigation of the mechanism of membrane permeabilization induced by medium‐length peptaibiotics. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
In this article, we review the relevant results obtained during almost 60 years of research on a specific aspect of stereochemistry, namely handedness preference and switches between right‐handed and left‐handed helical peptide structures generated by protein amino acids or appropriately designed, side‐chain modified analogs. In particular, we present and discuss here experimental and theoretical data on three categories of those screw‐sense issues: (i) right‐handed/left‐handed α‐helix transitions underwent by peptides rich in Asp, specific Asp β‐esters, and Asn; (ii) comparison of the preferred conformations adopted by helical host–guest peptide series, each characterized by an amino acid residue (e.g. Ile or its diastereomer aIle) endowed with two chiral centers in its chemical structure; and (iii) right‐handed (type I)/left‐handed (type II) poly‐(Pro)n helix transitions monitored for peptides rich in Pro itself or its analogs with a pyrrolidine ring substitution, particularly at the biologically important position 4. The unique modular and chiral properties of peptides, combined with their relatively easy synthesis, the chance to shape them into the desired conformation, and the enormous chemical diversity of their coded and non‐coded α‐amino acid building blocks, offer a huge opportunity to structural chemists for applications to bioscience and nanoscience problems. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
The standard collagen triple‐helix requires a perfect (Gly‐Xaa‐Yaa)n sequence, yet all nonfibrillar collagens contain interruptions in this tripeptide repeating pattern. Defining the structural consequences of disruptions in the sequence pattern may shed light on the biological role of sequence interruptions, which have been suggested to play a role in molecular flexibility, collagen degradation, and ligand binding. Previous studies on model peptides with 1‐ and 4‐residue interruptions showed a localized perturbation within the triple‐helix, and this work is extended to introduce natural collagen interruptions up to nine residue in length within a fixed (Gly‐Pro‐Hyp)n peptide context. All peptides in this set show decreases in triple‐helix content and stability, with greater conformational perturbations for the interruptions longer than five residue. The most stable and least perturbed structure is seen for the 5‐residue interruption peptide, whose sequence corresponds to a Gly to Ala missense mutation, such as those leading to collagen genetic diseases. The triple‐helix peptides containing 8‐ and 9‐residue interruptions exhibit a strong propensity for self‐association to fibrous structures. In addition, a small peptide modeling only the 9‐residue sequence within the interruption aggregates to form amyloid‐like fibrils with antiparallel β‐sheet structure. The 8‐ and 9‐residue interruption sequences studied here are predicted to have significant cross‐β aggregation potential, and a similar propensity is reported for ~10% of other naturally occurring interruptions. The presence of amyloidogenic sequences within or between triple‐helix domains may play a role in molecular association to normal tissue structures and could participate in observed interactions between collagen and amyloid.  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