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.
Polarized ir spectra of oriented films of α‐helical poly(l ‐alanine) (α‐PLA) have been obtained as a function of residual solvent dichloroacetic acid (DCA). The amide A, B, II, and V regions exhibit multiple bands whose structure depends on the residual DCA content, and those associated with the αI‐PLA structure have been identified. A calculation of the relevant cubic anharmonic force constants indicates that, contrary to previous assignments, the overtone of amide II(A) is in Fermi resonance with the NH stretch fundamental, whose unperturbed frequency we now find to be at 3314 cm−1, significantly higher than the previously suggested 3279 cm−1. The presence of a structure in addition to the standard αI‐PLA is indicated by our analysis. © 1999 John Wiley & Sons, Inc. Biopoly 49: 195–207, 1999  相似文献   

2.
The structural influence of a single Gly residue inserted into an Aib16 homooligomer was studied in the solid state by X‐ray crystallography. The peptides N3Aib8GlyAib8PheNH2 ( 1 ) and CbzPheAib8GlyAib8 ( 2 ) were found to adopt well‐defined helical structures, which are broadly 310 helical. Indeed, 2 is the longest crystallographic 310 helix thus far reported. However, in the region of the central Gly residue, a loosening of the 310 structure is observed in both peptides, with 1 clearly showing local adoption of an α‐helical structure in the region of residues 7–9. © 2010 Wiley Periodicals, Inc. Biopolymers 95: 62–69, 2011.  相似文献   

3.
One chiral L ‐valine (L ‐Val) was inserted into the C‐terminal position of achiral peptide segments constructed from α‐aminoisobutyric acid (Aib) and α,β‐dehydrophenylalanine (ΔZPhe) residues. The IR, 1H NMR and CD spectra indicated that the dominant conformations of the pentapeptide Boc‐Aib‐ΔPhe‐(Aib)2‐L ‐Val‐NH‐Bn (3) and the hexapeptide Boc‐Aib‐ΔPhe‐(Aib)3‐L ‐Val‐NH‐Bn (4) in solution were both right‐handed (P) 310‐helical structures. X‐ray crystallographic analyses of 3 and 4 revealed that only a right‐handed (P) 310‐helical structure was present in their crystalline states. The conformation of 4 was also studied by molecular‐mechanics calculations. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
Fourier transform infrared spectroscopy (FTIR) can be used for conformational analysis of peptides in a wide range of environments. Measurements can be performed in aqueous solution, organic solvents, detergent micelles as well as in phospholipid membranes. Information on the secondary structure of peptides can be derived from the analysis of the strong amide I band. Orientation of secondary structural elements within a lipid bilayer matrix can be determined by means of polarized attenuated total reflectance–FTIR spectroscopy. Hydrogen–deuterium exchange can be monitored by the analysis of the, amide II band. This review gives some example of peptide systems studied by FTIR spectroscopy. Studies on alamethicin and α-aminoisobutyric acid containing peptides have shown that FTIR spectroscopy is a sensitive tool for identifying 310-helical structures. Changes in the structure of the magainins upon interaction with charged lipids were detected using FTIR spectroscopy. Tachyplesin is an example of a β-sheet containing membrane active peptide. Polarized ir spectroscopy reveals that the antiparallel β-sheet structures of tachyplesin are oriented parallel to the membrane surface. Synthesis of peptides corresponding to functionally/structurally important regions of large proteins is becoming increasingly popular. FTIR spectroscopy has been used to analyze the structure of synthetic peptides corresponding to the ion-selective pore of the voltage-gated potassium channel. In biomembrane systems these peptides adopt a highly helical structure. Under conditions, where these peptides are aggregated the presence of some intermolecular β-sheet structure can also be detected. © 1994 John Wiley & Sons, Inc.  相似文献   

5.
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   

6.
The rotational strengths and the robustness values of amide‐I and amide‐II vibrational modes of For(AA)nNHMe (where AA is Val, Asn, Asp, or Cys, n = 1–5 for Val and Asn; n = 1 for Asp and Cys) model peptides with α‐helix and β‐sheet backbone conformations were computed by density functional methods. The robustness results verify empirical rules drawn from experiments and from computed rotational strengths linking amide‐I and amide‐II patterns in the vibrational circular dichroism (VCD) spectra of peptides with their backbone structures. For peptides with at least three residues (n ≥ 3) these characteristic patterns from coupled amide vibrational modes have robust signatures. For shorter peptide models many vibrational modes are nonrobust, and the robust modes can be dependent on the residues or on their side chain conformations in addition to backbone conformations. These robust VCD bands, however, provide information for the detailed structural analysis of these smaller systems. Chirality 27:625–634, 2015 © 2015 Wiley Periodicals, Inc.  相似文献   

7.
Melittin, the main hemolytic component of honeybee venom, is unfolded in an aqueous environment and folds into an α‐helical conformation in a lipid environment. Membrane fluidity is known to affect the activity and structure of melittin. By combining two structurally sensitive optical methods, circular dichroism (CD) and deep‐ultraviolet resonance Raman spectroscopy (dUVRR), we have identified distinct structural fluctuations in melittin correlated with increased and decreased 1,2‐dimyristoyl‐sn‐glycero‐3‐phosphocholine bilayer fluidities. CD spectra have reduced intensity at temperatures above 22°C and high concentrations of the cholesterol analog 5α‐cholestan‐3β‐ol indicating distortions in the α‐helical structure under these conditions. No increase in the amide S is observed in the temperature‐dependent dUVRR spectra, suggesting an increase in 310‐helical structure with increasing temperatures above 22°C. However, incorporation of 25 mol% 5α‐cholestan‐3β‐ol resulted in a small increase in the amide S intensity indicating partial unfolding of melittin. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 895–902, 2014.  相似文献   

8.
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   

9.
The right‐handed α‐helix is the dominant helical fold of α‐peptides, whereas the left‐handed 314‐helix is the dominant helical fold of β‐peptides. Using molecular dynamics simulations, the properties of α‐helical α‐peptides and 314‐helical β‐peptides with different C‐terminal protonation states and in the solvents water and methanol are compared. The observed energetic and entropic differences can be traced to differences in the polarity of the solvent‐accessible surface area and, in particular, the solute dipole moments, suggesting different reasons for their stability.  相似文献   

10.
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   

11.
Although the N‐terminal region in human apolipoprotein (apo) A‐I is thought to stabilize the lipid‐free structure of the protein, its role in lipid binding is unknown. Using synthetic fragment peptides, we examined the lipid‐binding properties of the first 43 residues (1–43) of apoA‐I in comparison with residues 44–65 and 220–241, which have strong lipid affinity in the molecule. Circular dichroism measurements demonstrated that peptides corresponding to each segment have potential propensity to form α‐helical structure in trifluoroethanol. Spectroscopic and thermodynamic measurements revealed that apoA‐I (1–43) peptide has the strong ability to bind to lipid vesicles and to form α‐helical structure comparable to apoA‐I (220–241) peptide. Substitution of Tyr‐18 located at the center of the most hydrophobic region in residues 1–43 with a helix‐breaking proline resulted in the impaired lipid binding, indicating that the α‐helical structure in this region is required to trigger the lipid binding. In contrast, apoA‐I (44–65) peptide exhibited a lower propensity to form α‐helical structure upon binding to lipid, and apoA‐I (44–65/S55P) peptide exhibited diminished, but not completely impaired, lipid binding, suggesting that the central region of residues 44–65 is not pivotally involved in the formation of the α‐helical structure and lipid binding. These results indicate that the most N‐terminal region of apoA‐I molecule, residues 1–43, contributes to the lipid interaction of apoA‐I through the hydrophobic helical residues. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
RNA cleaving conjugates have been prepared by attachment of oligodeoxyribonucleotide TTTT to peptides containing arginine, leucine, proline and serine residues. The highest activity was displayed by the conjugates containing peptides with alternating arginine and leucine residues (LR)4G‐amide. Ribonuclease activity of the conjugates pep‐T4 decreases in the order T4‐(LR)4G > T4‐(LR)2G > T4‐(LLRR)2G > T4‐(LR)2PRLRG > S2R3‐Hmda‐T4 ≥ R5 ≠ (LR)3. According to CD spectra, the free peptide (LR)4G‐amide in water solution at neutral pH and physiological ionic strength has no pronounced secondary structure whereas conjugated to oligonucleotide it acquires a folding similar to α‐helix.  相似文献   

13.
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   

14.
Linear cationic α‐helical antimicrobial peptides are promising chemotherapeutics. Most of them act by different mechanisms, making it difficult to microorganisms acquiring resistance. Decoralin is an example of antimicrobial peptide; it was described by Konno et al. and presented activity against microorganisms, but with pronounced hemolytic activity. We synthesized leucine‐substituted decoralin analogs designed based on important physicochemical properties, which depend on the maintenance of the amphiphilic α‐helical tendency of the native molecule. Peptides were synthesized, purified, and characterized, and the conformational studies were performed. The results indicated that the analogs presented both higher therapeutic indexes, but with antagonistic behavior. While [Leu]10‐Dec‐NH2 analog showed similar activity against different microorganisms (c.a. 0.4–0.8 μmol L?1), helical structuration, and some hemolytic activity, [Leu]8‐Dec‐NH2 analog did not tend to helical structure and presented antimicrobial activities two orders higher than the other two peptides analyzed. On the other hand, this analog showed to be the less hemolytic (MHC value = 50.0 μmol L?1). This approach provided insight for understanding the effects of the leucine substitution in the amphiphilic balance. They led to changes on the conformational tendency, which showed to be important for the mechanism of action and affecting antimicrobial and hemolytic activities. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
The synthetic, zwitterionic bacterial cell wall peptides—D -Gluγ-L-Lys, D -Gluγ-L-Lys-D -Ala, D -Gluγ-L-Lys-D -Ala-D -Ala, and L-Ala-D -Gluγ-L-Lys-D -Ala-D -Ala—have been investigated in the crystalline and aqueous solution state applying ir and Raman spectroscopy. Additionally, aqueous solutions of the tetra- and pentapeptide have been investigated by CD spectroscopic techniques. Apart from the dipeptide, whose spectral features were dominated by end-group vibrations, the corresponding ir and Raman active bands of the crystalline peptides in the amide and skeletal regions were found at similar wave numbers, thus suggesting an analogous three-dimensional structure of these compounds. Dominant amide A, I, II, and III bands near 3275, 1630, 1540, and 1220–1250 cm?1, respectively, in the ir are interpreted in favor of an intermolecularly hydrogen-bonded, β-like structure. The absence of any amide components near 1680–1690 cm?1, together with the presence of strong amide bands near 1630 cm?1, and weak bands near 1660 cm?1 in the ir, which, conversely, were found in the Raman spectra as weak and strong bands, but at corresponding wave numbers, is taken as strong evidence for the presence of the unusual, parallel-arranged β-structure. On the basis of comparative theoretical considerations, a parallel-arranged, “β-type ring” conformation [P. De Santis, S. Morosetti, and R. Rizzo (1974) Macromolecules 7 , 52–58] is hypothesized. The solubilized peptides exhibited distinct similarities with their crystalline counterparts in respect to frequency values and relative intensities of the corresponding ir and Raman-active amide I/I′ components, and of some Raman bands in the skeletal region. This is interpreted in terms of residual short-range order, persisting even in aqueous solution. We concluded that the peptides show a strong propensity to form hydrated, strongly associated aggregates in water. On the basis of amide I/I′ band positions, stable, intramolecular interactions via the amide groups are discarded for the solubilized peptides. Complementarily, the CD data obtained suggest the presence of weakly bent, “open-turn”-like structures for the tetra- and pentapeptide in aqueous solution.  相似文献   

16.
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   

17.
This work advances bottom‐up design of bioinspired materials built from peptide‐amphiphiles, which are a class of bioconjugates in which a biofunctional peptide is covalently attached to a hydrophobic moiety that drives self‐assembly in aqueous solution. Specifically, this work highlights the importance of peptide contour length in determining the equilibrium secondary structure of the peptide as well as the self‐assembled (i.e., micelle) geometry. Peptides used here repeat a seven‐amino acid sequence between one and four times to vary peptide contour length while maintaining similar peptide‐peptide interactions. Without a hydrophobic tail, these peptides all exhibit a combination of random coil and α‐helical structure. Upon self‐assembly in the crowded environment of a micellar corona, however, short peptides are prone to β‐sheet structure and cylindrical micelle geometry while longer peptides remain helical in spheroidal micelles. The transition to β‐sheets in short peptides is rapid, whereby amphiphiles first self‐assemble with α‐helical peptide structure, then transition to their equilibrium β‐sheet structure at a rate that depends on both temperature and ionic strength. These results identify peptide contour length as an important control over equilibrium peptide secondary structure and micelle geometry. Furthermore, the time‐dependent nature of the helix‐to‐sheet transition opens the door for shape‐changing bioinspired materials with tunable conversion rates. © 2013 Wiley Periodicals, Inc. Biopolymers 99: 573–581, 2013.  相似文献   

18.
Cationic antimicrobial peptides have attracted increasing attention as a novel class of antibiotics to treat infectious diseases caused by pathogenic bacteria. However, susceptibility to protease is a shortcoming in their development. Cyclization is one approach to increase the proteolytic resistance of peptides. Therefore, to improve the proteolytic resistance of Polybia‐MPI, we have synthesized the MPI cyclic analogs C‐MPI‐1 (i‐to‐i+4) and C‐MPI‐2 (i‐to‐i+6) by copper(I)‐catalyzed azide–alkyne cycloaddition. Compared with MPI, C‐MPI‐1 displayed sustained antimicrobial activity and had enhanced anti‐trypsin resistance, while C‐MPI‐2 displayed no antimicrobial activity. The relationship between peptide structure and bioactivity was further investigated by probing the secondary structure of the peptides by circular dichroism. This showed that C‐MPI‐1 adopted an α‐helical structure in aqueous solution and, interestingly, had increased α‐helical conformation in 30 mM sodium dodecyl sulfate and 50% trifluoroethyl alcohol compared with MPI. C‐MPI‐2 that was not α‐helical in structure, suggesting that the propensity for α‐helix conformation may play an important role in cyclic peptide design. In addition, scanning electron microscopy, propidium iodide uptake, and membrane permeabilization assays indicated that MPI and the optimized analog C‐MPI‐1 had membrane‐active action modes, indicating that the peptides would not be susceptible to conventional resistance mechanisms. Our study provides additional insight into the influence of intramolecular cyclization at various positions on peptide structure and biological activity. In conclusion, the design and synthesis of cyclic analogs via click chemistry offer a new strategy for the development of stable antimicrobial agents. Copyright © 2017 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
The pentapeptide Boc-Val-ΔPhe-Gly-ΔPhe-Val-OMe, containing two dehydro-phenylalanine (ΔPhe) residues, has been synthesized and its structure investigated. In the crystalline state, the molecule adopts a right-handed 310-helical conformation stabilized by two intramolecular hydrogen bonds between CO of Val1 and NH of ΔPhe4, and between CO of ΔPhe2 and NH of Val5, respectively. NMR measurements are consistent with the presence of 310-helical structures also in acetonitrile and dimethylsulphoxide solution: the distances between backbone protons estimated from NOE connectivities are in overall agreement with those observed in the solid state; the chemical shifts of the amide protons show the smaller temperature coefficients for the NHs that in solid state are involved in intramolecular hydrogen bonds. The CD spectra in acetonitrile, chloroform, methanol and dimethylsulphoxide display exciton couplets of bands corresponding to the ΔPhe electronic transition at 280nm; the sign of the bands is consistent with the presence of helical structures having a prevalent left-handed screw sense. Addition of 1,1,1,3,3,3-hexafluoro- propan-2-ol gives rise to the gradual appearance of a couplet of opposite sign, suggesting the helix reversal from left-handed sense to right-handed sense. The conformational behaviour is discussed on the basis of the specific sequence of the peptide.  相似文献   

20.
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   

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