首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The polyproline type II (PPII) helix is a prevalent conformation in both folded and unfolded proteins, and is known to play important roles in a wide variety of biological processes. Polyproline itself can also form a type I (PPI) helix, which has a disparate conformation. Here, we use derivatives of polyproline, (Pro)10, (Hyp)10, (Flp)10, and (flp)10, where Hyp is (2S,4R)-4-hydroxyproline, Flp is (2S,4R)-4-fluoroproline, and flp is (2S,4S)-4-fluoroproline, to probe for a stereoelectronic effect on the conformation of polyproline. Circular dichroism spectral analyses show that 4R electron-with-drawing substituents stabilize a PPII helix relative to a PPI helix, even in a solvent that favors the PPI conformation, such as n-propanol. The stereochemistry at C4 ordains the relative stability of PPI and PPII helices, as (flp)10 forms a mixture of PPI and PPII helices in water and a PPI helix in n-propanol. The conformational preferences of (Pro)10 are intermediate between those of (Hyp)10/(Flp)10 and (flp)10. Interestingly, PPI helices of (flp)10 exhibit cold denaturation in n-propanol with a value of T(s) near 70 degrees C. Together, these data show that stereoelectronic effects can have a substantial impact on polyproline conformation and provide a rational means to stabilize a PPI or PPII helix.  相似文献   

2.
The polyproline‐II helix is the most extended naturally occurring helical structure and is widely present in polar, exposed stretches and “unstructured” denatured regions of polypeptides. Can it be hydrophobic? In this study, we address this question using oligomeric peptides formed by a hydrophobic proline analogue, (2S,3aS,7aS)‐octahydroindole‐2‐carboxylic acid (Oic). Previously, we found the molecular principles underlying the structural stability of the polyproline‐II conformation in these oligomers, whereas the hydrophobicity of the peptide constructs remains to be examined. Therefore, we investigated the octan‐1‐ol/water partitioning and inclusion in detergent micelles of the oligo‐Oic peptides. The results showed that the hydrophobicity is remarkably enhanced in longer oligomeric sequences, and the oligo‐Oic peptides with 3 to 4 residues and higher are specific towards hydrophobic environments. This contrasts significantly to the parent oligoproline peptides, which were moderately hydrophilic. With these findings, we have demonstrated that the polyproline‐II structure is compatible with nonpolar media, whereas additional manipulations of the terminal functionalities feature solubility in extremely nonpolar solvents such as hexane.  相似文献   

3.
The folding of proteins is directed by a variety of interactions, including hydrogen bonding, electrostatics, van der Waals' interactions, and the hydrophobic effect. We have argued previously that an n→π* interaction between carbonyl groups be added to this list. In an n→π* interaction, the lone pair (n) of one carbonyl oxygen overlaps with the π* antibonding orbital of another carbonyl group. The tendency of backbone carbonyl groups in proteins to engage in this interaction has consequences for the structures of folded proteins that we unveil herein. First, we employ density functional theory to demonstrate that the n→π* interaction causes the carbonyl carbon to deviate from planarity. Then, we detect this signature of the n→π* interaction in high‐resolution structures of proteins. Finally, we demonstrate through natural population analysis that the n→π* interaction causes polarization of the electron density in carbonyl groups and detect that polarization in the electron density map of cholesterol oxidase, further validating the existence of n→π* interactions. We conclude that the n→π* interaction is operative in folded proteins.  相似文献   

4.
The oxygen of a peptide bond has two lone pairs of electrons. One of these lone pairs is poised to interact with the electron-deficient carbon of the subsequent peptide bond in the chain. Any partial covalency that results from this n→π* interaction should induce pyramidalization of the carbon (C'(i)) toward the oxygen (O(i-1)). We searched for such pyramidalization in 14 peptides that contain both α- and β-amino acid residues and that assume a helical structure. We found that the α-amino acid residues, which adopt the main chain dihedral angles of an α-helix, display dramatic pyramidalization but the β-amino acid residues do not. Thus, we conclude that O(i-1) and C'(i) are linked by a partial covalent bond in α-helices. This finding has important ramifications for the folding and conformational stability of α-helices in isolation and in proteins.  相似文献   

5.
Collagens have long been believed to adopt a triple‐stranded molecular structure with a 10/3 symmetry (ten triplet units in three turns) and an axial repeat of 29 Å. This belief even persisted after an alternative structure with a 7/2 symmetry (seven triplet units in two turns) with an axial repeat of 20 Å had been proposed. The uncertainty regarding the helical symmetry of collagens is attributed to inadequate X‐ray fiber diffraction data. Therefore, for better understanding of the collagen helix, single‐crystal analyses of peptides with simplified characteristic amino acid sequences and similar compositions to collagens have long been awaited. Here we report the crystal structure of (Gly‐Pro‐Hyp)9 peptide at a resolution of 1.45 Å. The repeating unit of this peptide, Gly‐Pro‐Hyp, is the most typical sequence present in collagens, and it has been used as a basic repeating unit in fiber diffraction analyses of collagen. The (Gly‐Pro‐Hyp)9 peptide adopts a triple‐stranded structure with an average helical symmetry close to the ideal 7/2 helical model for collagen. This observation strongly suggests that the average molecular structure of collagen is not the accepted Rich and Crick 10/3 helical model but is a 7/2 helical conformation. © 2012 Wiley Periodicals, Inc. Biopolymers 97: 607–616, 2012.  相似文献   

6.
There has been growing interest in polyproline type II (PPII) helices since PPII helices have been found in folded and unfolded proteins and involved in a variety of biological activities. Polyproline can also form type I helices (PPI) which are very different from PPII conformation and only exist in certain organic solvents. Recent studies have shown that stereoelectronic effects play a critical role in stabilizing a PPI or PPII helix. Here, we have synthesized a series of host–guest peptides with an electron‐withdrawing substituent at the 4R or 4S position of proline and used a kinetic approach to further explore stereoelectronic effects on the transition barrier of the interconversion between PPI and PPII conformations. Time‐dependent circular dichroism measurements revealed that the rates of PPII → PPI conversion were reduced upon incorporating the hydroxyl‐, fluoro‐, and methoxy‐groups at the 4R position while the rates would be increased if these substituents were at the 4S position. We quantified the changes in transition free energy by comparing their rate constants. (4R,2S)‐4‐Fluoroproline and (4S,2S)‐4‐fluoroproline have the largest effect on the transition energy barrier for PPII → PPI conversion. Our results provide important insights into the role of stereoelectronic effects on the PPII → PPI transition state barrier, which has not been reported in past thermodynamic studies.  相似文献   

7.
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   

8.
Tat cell-penetrating peptide (GRKKRRQRRRPPQG) is able to translocate and carry molecules across cell membranes. Using CD spectroscopy the conformation of this synthetic peptide was studied in aqueous and membrane-mimicking, micellar SDS solutions at different temperatures. The CD spectrum of the Tat cell-penetrating peptide in SDS micellar solution was virtually unchanged from that in aqueous solution, and at low temperature it was close to that of a poly(proline) II helix.  相似文献   

9.
In the selection or design of antimicrobial peptides, the key role played by cationic amino acids and chain length on the inhibitory potency and specificity is not clear. A fundamental study was conducted using chemically synthesized homopeptides of l ‐Lys and l ‐Arg ranging from 7 to 14 residues. Their effect on growth inhibition was evaluated over a wide range of Gram‐positive bacteria at different levels of concentration. Interestingly, at lower concentrations (10 μM), Lys homopeptides with odd number of residues, especially with 11 residues, showed a broader inhibitory activity than those with even number of residues. At higher peptide concentrations (>20 μM), the inhibitory activity of Lys homopeptides was directly related to the number of residues in the chain. In contrast, Arg homopeptides, at lower concentrations, did not exhibit a defined pattern of bacterial inhibition related to the number of residues; however, at higher concentrations (>20 μM), the inhibitory effects were more pronounced. Lys homopeptides at concentrations up to 300 μM showed a remarkably lower toxicity against CHSE‐214 cells. Arg homopeptides exhibited negligible cytotoxicity up to chain length of 11 residues at concentrations lower than 100 μM, but an abrupt increase in toxicity resulted when the peptide chain length reached 12 amino acid residues and higher concentrations. All synthesized homopeptides displayed characteristic polyproline II helix conformation in both buffer and liposomes, as shown by CD spectroscopy. This result suggests that short Lys homopeptides with an odd number of residues (9 and 11) have a broad spectrum of activity against Gram‐positive bacterial cells compared with Arg homopeptides, which in turn showed a considerably higher selectivity toward those cells. By investigating the differences between Lys and Arg homopeptides, this study contributes to the understanding of their mechanism of growth inhibition and selectivity. Thus, it provides further guidelines for a rational design of short antimicrobial peptides. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
The proline-rich N-terminal domain of gamma-zein has been reported in relevant processes, which include its ability to cross the cell membranes. Evidences indicate that synthetic hexapeptide (PPPVHL), naturally found in N-terminal portion of gamma-zein, can adopt the polyproline II (PPII) conformation in aqueous solution. The secondary structure of gamma-zein in maize protein bodies had been analyzed by solid state Fourier transform infrared and nuclear magnetic resonance spectroscopies. However, it was not possible to measure PPII content in physiological environment since the beta-sheet and PPII signals overlap in both solid state techniques. Here, the secondary structure of gamma-zein has been analyzed by circular dichroism in SDS aqueous solution with and without ditiothreitol (DTT), and in 60% of 2-propanol and water with DTT. The results show that gamma-zein has high helical content in all solutions. The PPII conformation was present at about 7% only in water/DTT solution.  相似文献   

11.
The protein folding process is often in vitro rate‐limited by slow cis‐trans proline isomerization steps. Importantly, the rate of this process in vivo is accelerated by prolyl isomerases (PPIases). The archetypal PPIase is the human cyclophilin 18 (Cyp18 or CypA), and Arg 55 has been demonstrated to play a crucial role when studying short peptide substrates in the catalytic action of Cyp18 by stabilizing the transition state of isomerization. However, in this study we show that a R55A mutant of Cyp18 is as efficient as the wild type to accelerate the refolding reaction of human carbonic anhydrase II (HCA II). Thus, it is evident that the active‐site located Arg 55 is not required for catalysis of the rate‐limiting prolyl cis‐trans isomerization steps during the folding of a protein substrate as HCA II. Nevertheless, catalysis of cis‐trans proline isomerization in HCA II occurs in the active‐site of Cyp18, since binding of the inhibitor cyclosporin A abolishes rate acceleration of the refolding reaction. Obviously, the catalytic mechanisms of Cyp18 can differ when acting upon a simple model peptide, four residues long, with easily accessible Pro residues compared with a large protein molecule undergoing folding with partly or completely buried Pro residues. In the latter case, the isomerization kinetics are significantly slower and simpler mechanistic factors such as desolvation and/or strain might operate during folding‐assisted catalysis, since binding to the hydrophobic active site is still a prerequisite for catalysis.  相似文献   

12.
Chellgren BW  Creamer TP 《Proteins》2006,62(2):411-420
Loss of conformational entropy is one of the primary factors opposing protein folding. Both the backbone and side-chain of each residue in a protein will have their freedom of motion restricted in the final folded structure. The type of secondary structure of which a residue is part will have a significant impact on how much side-chain entropy is lost. Side-chain conformational entropies have previously been determined for folded proteins, simple models of unfolded proteins, alpha-helices, and a dipeptide model for beta-strands, but not for polyproline II (PII) helices. In this work, we present side-chain conformational estimates for the three regular secondary structure types: alpha-helices, beta-strands, and PII helices. Entropies are estimated from Monte Carlo computer simulations. Beta-strands are modeled as two structures, parallel and antiparallel beta-strands. Our data indicate that restraining a residue to the PII helix or antiparallel beta-strand conformations results in side-chain entropies equal to or higher than those obtained by restraining residues to the parallel beta-strand conformation. Side-chains in the alpha-helix conformation have the lowest side-chain entropies. The observation that extended structures retain the most side-chain entropy suggests that such structures would be entropically favored in unfolded proteins under folding conditions. Our data indicate that the PII helix conformation would be somewhat favored over beta-strand conformations, with antiparallel beta-strand favored over parallel. Notably, our data imply that, under some circumstances, residues may gain side-chain entropy upon folding. Implications of our findings for protein folding and unfolded states are discussed.  相似文献   

13.
Proline is an amino acid with a unique cyclic structure that facilitates the folding of many proteins, but also impedes the rate of peptide bond formation by the ribosome. As a ribosome substrate, proline reacts markedly slower when compared with other amino acids both as a donor and as an acceptor of the nascent peptide. Furthermore, synthesis of peptides with consecutive proline residues triggers ribosome stalling. Here, we report crystal structures of the eukaryotic ribosome bound to analogs of mono‐ and diprolyl‐tRNAs. These structures provide a high‐resolution insight into unique properties of proline as a ribosome substrate. They show that the cyclic structure of proline residue prevents proline positioning in the amino acid binding pocket and affects the nascent peptide chain position in the ribosomal peptide exit tunnel. These observations extend current knowledge of the protein synthesis mechanism. They also revise an old dogma that amino acids bind the ribosomal active site in a uniform way by showing that proline has a binding mode distinct from other amino acids.  相似文献   

14.
The solid phase synthesis of a 59 amino acid human salivary protein IB7 has been accomplished using Fmoc strategy. Because the protein contains 25 proline, 13 glycine and 9 glutamine residues the coupling time, piperidine delivery and acetic anhydride reaction time were increased. Yield after HPLC purification was 35%. Circular dichroism studies revealed that about one third of IB7 residues adopted a type II helix secondary structure, as found in collagen helices. The rest of the sequence adopts a random coil secondary structure.  相似文献   

15.
The effect of choline chloride on the conformational dynamics of the 11‐mer repeat unit P1LEA‐22 of group 3 Late Embryogenesis Abundant (G3LEA) proteins was studied. Circular dichroism data of aqueous solutions of P1LEA‐22 revealed that the peptide favors a polyproline II (PPII) helix structure at low temperature, with increasing temperature promoting a gain of unstructured conformations. Furthermore, increases in sample FeCl3 or choline chloride concentrations causes a gain in PPII helical structure at low temperature. The potential role of PPII structure in intrinsically disordered and G3LEA proteins is discussed, including its ability to easily access other secondary structural conformations such as α‐helix and β‐sheet, which have been observed for dehydrated G3LEA proteins. The observed effect of FeCl3 and choline chloride salts on P1LEA‐22 suggests favorable cation interactions with the PPII helix, supporting ion sequestration as a G3LEA protein function. As choline chloride is suggested to improve salt tolerance and protect cell membrane in plants at low temperature, our results support adoption of the PPII structure as a possible damage‐preventing measure of Late Embryogenesis Abundant proteins.  相似文献   

16.
Vibrational Raman optical activity (ROA), measured as a small difference in the intensity of Raman scattering from chiral molecules in right- and left-circularly polarized incident light, or as the intensity of a small circularly polarized component in the scattered light, is a powerful probe of the aqueous solution structure of proteins. The large number of structure-sensitive bands in protein ROA spectra makes multivariate analysis techniques such as nonlinear mapping (NLM) especially favorable for determining structural relationships between different proteins. We have previously used NLM to map a large dataset of peptide, protein, and virus ROA spectra into a readily visualizable two-dimensional space in which points close to or distant from each other, respectively, represent similar or dissimilar structures. As well as folded proteins, our dataset contains ROA spectra from many natively unfolded proteins, proteins containing both folded and unfolded domains, denatured partially structured molten globule and reduced protein states, together with folded proteins containing little or no alpha-helix or beta-sheet. In this article, the relative positions of these systems in the NLM plot are used to obtain information about any residual structure that they may contain. The striking differences between the structural propensities of proteins that are unfolded in their native states and those that are unfolded due to denaturation may be responsible for their often very different behavior, especially with regard to aggregation. An ab initio simulation of the Raman and ROA spectra of an alanine oligopeptide in the poly(L-proline) II-helical conformation confirms previous suggestions that this conformation is a significant structural element in disordered peptides and natively unfolded proteins. The use of ROA to identify and characterize proteins containing significant amounts of unfolded structure will, inter alia, be valuable in structural genomics/proteomics since unfolded sequences often inhibit crystallization.  相似文献   

17.
The tumor suppresser protein p53 has been called the “guardian of the genome.” DNA damage induces p53 to either halt the cell cycle, allowing for repair, or initiate apoptosis. P53 is mutated in over 50% of human tumors and it has been proposed that many tumorigenic mutations are deleterious to p53 because they induce local unfolding. To explore this hypothesis, peptide models have been developed to study tumorigenic mutations in the H2 helix of the p53 core domain. This helix is rich with charged residues and is a key component of the DNA binding region. A 16‐residue peptide corresponding to the H2 wild‐type sequence extended with an Ala‐rich C‐terminus was synthesized and studied by 1H‐nmr (500 MHz) and CD. The nmr studies demonstrate that this peptide adopts helical structure in solution. Six additional peptides corresponding to subtle tumorigenic mutations were synthesized and CD was used to assess the relative stability of these “mutant analogues.” All six mutations studied are destabilizing relative to the wild type, with ΔΔG values in the range of 0.26 to 1.35 kcal mol−1. Surprisingly, substitution of Asp 281 with Ala resulted in a peptide with the greatest destabilization even though Ala possesses the largest helix propensity of the common 20 amino acids. Because this helix appears to be stabilized mainly by local electrostatics, we conclude that its structure is susceptible to even the most conservative mutations. These results provide support for the hypothesis that tumorigenic mutations induce local unfolding of p53. © 1999 John Wiley & Sons, Inc. Biopoly 49: 215–224, 1999  相似文献   

18.
Rheumatoid arthritis is an autoimmune disease associated with the recognition of self proteins secluded in arthritic joints. We generated transgenic rice seeds expressing three types of altered peptide ligands (APL) and the T cell epitope of type II collagen (CII256–271). When these transgenic rice and non-transgenic rice seeds were orally administrated to DBA/1?J mice once a day for 14?days, followed by immunization with CII, the clinical score of collagen-induced arthritis (CIA) was reduced and inflammation and erosion in the joints were prevented in mice fed APL7 transgenic rice only. IL-10 production against the CII antigen significantly increased in the splenocytes and iLN of CIA mice immunized with the CII antigen, whereas IFN-γ, IL-17, and IL-2 levels were not altered. These results suggest that IL-10-mediated immune suppression is involved in the prophylactic effects caused by transgenic rice expressing APL7.  相似文献   

19.
Co‐administration of beta‐lactam antibiotics and beta‐lactamase inhibitors has been a favored treatment strategy against beta‐lactamase‐mediated bacterial antibiotic resistance, but the emergence of beta‐lactamases resistant to current inhibitors necessitates the discovery of novel non‐beta‐lactam inhibitors. Peptides derived from the Ala46–Tyr51 region of the beta‐lactamase inhibitor protein are considered as potent inhibitors of beta‐lactamase; unfortunately, peptide delivery into the cell limits their potential. The properties of cell‐penetrating peptides could guide the design of beta‐lactamase inhibitory peptides. Here, our goal is to modify the peptide with the sequence RRGHYY that possesses beta‐lactamase inhibitory activity under in vitro conditions. Inspired by the work on the cell‐penetrating peptide pVEC, our approach involved the addition of the N‐terminal hydrophobic residues, LLIIL, from pVEC to the inhibitor peptide to build a chimera. These residues have been reported to be critical in the uptake of pVEC. We tested the potential of RRGHYY and its chimeric derivative as a beta‐lactamase inhibitory peptide on Escherichia coli cells and compared the results with the action of the antimicrobial peptide melittin, the beta‐lactam antibiotic ampicillin, and the beta‐lactamase inhibitor potassium clavulanate to get mechanistic details on their action. Our results show that the addition of LLIIL to the N‐terminus of the beta‐lactamase inhibitory peptide RRGHYY increases its membrane permeabilizing potential. Interestingly, the addition of this short stretch of hydrophobic residues also modified the inhibitory peptide such that it acquired antimicrobial property. We propose that addition of the hydrophobic LLIIL residues to the peptide N‐terminus offers a promising strategy to design novel antimicrobial peptides in the battle against antibiotic resistance. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
The molecular mechanism by which HFIP stabilizes the alpha-helical structure of peptides is not well understood. In the present study, we use melittin as a model to gain insight into the details of the atomistic interactions of HFIP with the peptide. We have performed extensive comparative molecular dynamics simulations (up to 100 nsec) in the absence and in the presence of HFIP. In agreement with recent NMR experiments, the simulations show rapid loss of tertiary structure in water at pH 2 but much higher helicity in 35% HFIP. The MD simulations also indicate that melittin adopts a highly dynamic global structure in 35% HFIP solution with two alpha-helical segments sampling a wide range of angular orientations. The analysis of the HFIP distribution shows the tendency of HFIP to aggregate around the peptide, increasing the local cosolvent concentration to more than two times that in the bulk concentration. The correlation of local peptide structure with HFIP coating suggests that displacement of water at the peptide surface is the main contribution of HFIP in stabilizing the secondary structure of melittin. Finally, a stabilizing effect promoted by the presence of counter-ions was also observed in the simulations.  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