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.
Z-Dehydrophenylalanine (ΔzPhe) possessing four oligopeptides, Boc-(L -Ala-ΔzPhe-Aib)n-OCH3 (n = 1–4: Boc, t-butoxycarbonyl; Aib, α-aminoisobutyric acid), were synthesized, and their solution conformations were investigated by 1H-nmr, ir, uv, and CD spectroscopy and theoretical CD calculation. 1H-nmr (the solvent accessibility of NH groups) and ir studies indicated that all the NH groups except for those belonging to the N-terminal L -Ala-ΔzPhe moiety participate in intramolecular hydrogen bonding in chloroform. This suggests that the peptides n = 2–4 have a 4 → 1 hydrogen-bonding pattern characteristic of 310-helical structures. The uv spectra of all these peptides recorded in chloroform and in trimethyl phosphate showed an intense maximum around 276 nm assigned to the ΔzPhe chromophores. The corresponding CD spectra of the peptides n = 2–4 showed exciton couplets with a negative peak at longer wavelengths, whereas that of the peptide n = 1 showed only weak signals. Theoretical CD spectra were calculated for the peptides n = 2–4 of several helical conformations, on the basis of exciton chirality method. This calculation indicated that the three peptides form a helical conformation deviating from the perfect 310-helix that contains three residues per turn, and that their side chains of Δz Phe residues are arranged regularly along the helix. The center-to-center distance between the nearest phenyl pair(s) was estimated to be ~ 5.5 Å. The chemical shifts of the ΔzPhe side-chain protons (Hβ and aromatic H) for the peptides n = 2–4 indicated anisotropic shielding effect of neighboring phenyl group(s); the effect also supports a regular arrangement of the Δz Phe side chains along the helical axis. © 1993 John Wiley & Sons, Inc.  相似文献   

2.
To understand the terminal effect of chiral residue for determining a helical screw sense, we adopted five kinds of peptides IV containing N‐ and/or C‐terminal chiral Leu residue(s): Boc–L ‐Leu–(Aib–ΔPhe)2–Aib–OMe ( I ), Boc–(Aib–ΔPhe)2–L ‐Leu–OMe ( II ), Boc–L ‐Leu–(Aib–ΔPhe)2–L ‐Leu–OMe ( III ), Boc–D ‐Leu–(Aib–ΔPhe)2–L ‐Leu–OMe ( IV ), and Boc–D ‐Leu–(Aib–ΔPhe)2–Aib–OMe ( V ). The segment –(Aib–ΔPhe)2– was used for a backbone composed of two “enantiomeric” (left‐/right‐handed) helices. Actually, this could be confirmed by 1H‐nmr [nuclear Overhauser effect (NOE) and solvent accessibility of NH resonances] and CD spectroscopy on Boc–(Aib–ΔPhe)2–Aib–OMe, which took a left‐/right‐handed 310‐helix. Peptides IV were also found to take 310‐type helical conformations in CDCl3, from difference NOE measurement and solvent accessibility of NH resonances. Chloroform, acetonitrile, methanol, and tetrahydrofuran were used for CD measurement. The CD spectra of peptides IIII in all solvents showed marked exciton couplets with a positive peak at longer wavelengths, indicating that their main chains prefer a left‐handed screw sense over a right‐handed one. Peptide V in all solvents showed exciton couplets with a negative peak at longer wavelengths, indicating it prefers a right‐handed screw sense. Peptide IV in chloroform showed a nonsplit type CD pattern having only a small negative signal around 280 nm, meaning that left‐ and right‐handed helices should exist with almost the same content. In the other solvents, peptide IV showed exciton couplets with a negative peak at longer wavelengths, corresponding to a right‐handed screw sense. From conformational energy calculation and the above 1H‐nmr studies, an N‐ or C‐terminal L ‐Leu residue in the lowest energy left‐handed 310‐helical conformation was found to take an irregular conformation that deviates from a left‐handed helix. The positional effect of the L ‐residue on helical screw sense was discussed based on CD data of peptides IV and of Boc–(L ‐Leu–ΔPhe)n–L ‐Leu–OMe (n = 2 and 3). © 1999 John Wiley & Sons, Inc. Biopoly 49: 551–564, 1999  相似文献   

3.
Two series of peptides containing L -phenylalanine, Nps-(L -Phe-L -Phe-Gly)n-OEt (n = 1–6) and Nps-(L -Phe-L -Leu-Gly)n-OEt (n = 1–7), were prepared by the fragment-condensation method using the tripeptide N-hydroxysuccinimide esters. Conformational characterization of these peptides in the solid state was performed by ir spectroscopy and x-ray powder diffraction measurement. The peptides Nps-(L -Phe-L -Phe-Gly)n-OEt take the β-structure, but the pentadecapeptide and higher peptides of Nps-(L -Phe-L -Leu-Gly)n-OEt form the α-helix, although the lower homologs take the β-structure.  相似文献   

4.
The solid state conformations of cyclo[Gly–Proψ[CH2S]Gly–D –Phe–Pro] and cyclo[Gly–Proψ[CH2–(S)–SO]Gly–D –Phe–Pro] have been characterized by X-ray diffraction analysis. Crystals of the sulfide trihydrate are orthorhombic, P212121, with a = 10.156(3) Å, b = 11.704(3) Å, c = 21.913(4) Å, and Z = 4. Crystals of the sulfoxide are monoclinic, P21, with a = 10.662(1) Å, b = 8.552(3) Å, c = 12.947(2) Å, β = 94.28(2), and Z = 2. Unlike their all-amide parent, which adopts an all-trans backbone conformation and a type II β-turn encompassing Gly-Pro-Gly-D -Phe, both of these peptides contain a cis Gly1-Pro2 bond and form a novel turn structure, i.e., a type II′ β-turn consisting of Gly–D –Phe–Pro–Gly. The turn structure in each of these peptides is stabilized by an intramolecular H bond between the carbonyl oxygen of Gly1 and the amide proton of D -Phe4. In the cyclic sulfoxide, the sulfinyl group is not involved in H bonding despite its strong potential as a hydrogen-bond acceptor. The crystal structure made it possible to establish the absolute configuration of the sulfinyl group in this peptide. The two crystal structures also helped identify a type II′ β-turn in the DMSO-d6 solution conformers of these peptides. © 1993 John Wiley & Sons, Inc.  相似文献   

5.
In order to elucidate the relationship between bitter taste and chemical structure in peptides, various kinds of model bitter peptides containing arginine, proline and phenylalanine were synthesized, and the contribution of the individual amino acids to the bitter taste was made clear. It was confirmed that, in order to strengthen the bitterness in di- and tripeptides, the hydrophobic amino acid needs to be located at the C-terminal and, conversely, the basic amino acid should be located at the N-terminal Furthermore, a strong bitter taste was observed when arginine was contiguous to proline such as Arg-Pro, Gly-Arg-Pro and Arg-Pro-Gly. A synergistic effect for bitter taste was observed in the peptides whose structure is (Arg)l-(Pro)m-(Phe)n (l=1, 2; m, n = 1 ~ 3) by increasing the number of amino acids. Among them, the octapeptide (Arg-Arg-Pro-Pro-Pro-Phe-Phe-Phe) possessed an extremely bitter taste with its threshold value of 0.002 mm and was found to be the most bitter among the peptides.  相似文献   

6.
This study designs a prediction model to differentiate pasteurized milk from heated extended shelf life (ESL) milk based on milk peptides. For this purpose, quantitative peptide profiles of a training set of commercial samples including pasteurized (n = 20), pasteurized‐ESL (n = 13), and heated‐ESL (n = 16) milk are recorded by matrix‐assisted laser desorption/ionization time‐of‐flight mass spectrometry (MALDI‐TOF‐MS). Seven peptides are selected as putative markers, and cutoff levels and performance measures of each marker are defined by receiver operating characteristic (ROC) analysis. The accuracy of these peptides in the training set range between 71% and 90%. A prediction model is established based on the combined cutoff levels and evaluated by an independent blind test set. The processing method of 19 out of 20 unknown milk samples is predicted correctly achieving 95% accuracy. Five peptides of the prediction model are identified as αS1‐casein182–199 (m/z 2014.0), αS1‐casein180–199 (m/z 2216.1), αS1‐casein1–24 (m/z 2910.6), β‐casein108–125 (m/z 2126.0), and β‐casein106–125 (m/z 2391.2) indicating thermal release and the action of plasmin and cathepsins. Thus, the present study demonstrates that the milk peptide profile reflects even minor differences in production parameters.  相似文献   

7.
Fifteen new peptide derivatives of ?-aminocaproic acid (EACA) containing the known fragment –Ala–Phe–Lys– with an affinity for plasmin were synthesised in the present study. The synthesis was carried out a solid phase. The following compounds were synthesised: H–Phe–Lys–EACA–X, H–d-Ala–Phe–Lys–EACA–X, H–Ala–Phe–Lys–EACA–X, H–d-Ala–Phe–EACA–X and H–Ala–Phe–EACA–X, where X = OH, NH2 and NH–(CH2)5–NH2. All peptides, except for those containing the sequence H–Ala–Phe–EACA–X, displayed higher inhibitory activity against plasmin than EACA. The most active and selective inhibitor of plasmin was the compound H–d-Ala–Phe–Lys–EACA–NH2 which inhibited the amidolytic activity of plasmin (IC50 = 0.02 mM), with the antifibrinolytic activity weaker than EACA. The resulting peptides did not affect the viability of fibroblast cells, colon cancer cell line DLD-1, breast MCF-7 and MDA-MB-231 cell lines.  相似文献   

8.
A series of N-acetyl-l-phenylalanyl peptides of general formula Ac-Phe-(Gly)n-NH2 (n = 0–2) has been synthesized to study the effect of leaving group chain length on the efficiency of chymotrypsin Aα amidase and peptidase activities. The effect upon catalysis of hydrophobic side chains on the leaving group was investigated using similar substrates with one of the glycine residues selectively substituted by an alanine residue as in AcPheAlaNH2, AcPheAlaGlyNH2, and AcPheGlyAlaNH2. Values of kcat and Km have been obtained from kinetic measurements at pH 8.00 and 25 °C. The results are shown to be consistent with binding schemes postulated from published model building studies. The catalytic reactions were studied over a range of temperature (15–35 °C) and in each case the Arrhenius law was obeyed. It was thus possible to obtain meaningful values for the thermodynamic functions of activation for the acylation step of the catalytic reaction. The results are shown to confirm the findings of postulated binding schemes but indicate that conclusions drawn from kinetic measurements at a single temperature may sometimes be misleading.  相似文献   

9.
Elastin, a core protein of the elastic fibers, exhibits the coacervation (temperature‐dependent reversible association/dissociation) under physiological conditions. Because of this characteristic, elastin and elastin‐derived peptides have been considered to be useful as base materials for developing various biomedical products, skin substitutes, synthetic vascular grafts, and drug delivery systems. Although elastin‐derived polypeptide (Val‐Pro‐Gly‐Val‐Gly)n also has been known to demonstrate coacervation property, a sufficiently high (VPGVG)n repetition number (n > 40) is required for coacervation. In the present study, a series of elastin‐derived peptide (Phe‐Pro‐Gly‐Val‐Gly)5 dimers possessing high coacervation potential were newly developed. These novel dimeric peptides exhibited coacervation at significantly lower concentrations and temperatures than the commonly used elastin‐derived peptide analogs; this result suggests that the coacervation ability of the peptides is enhanced by dimerization. Circular dichroism (CD) measurements indicate that the dimers undergo similar temperature‐dependent and reversible conformational changes when coacervation occurs. The molecular dynamics calculation results reveal that the sheet‐turn‐sheet motif involving a type II β‐turn‐like structure commonly observed among the dimers and caused formation of globular conformation of them. These synthesized peptide dimers may be useful not only as model peptides for structural analysis of elastin and elastin‐derived peptides, but also as base materials for developing various temperature‐sensitive biomedical and industrial products. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
Abstract: Calcitonin gene-related peptide (CGRP), a 37-amino-acid peptide, is a member of a small family of peptides including amylin or islet amyloid polypeptide and salmon calcitonin. These related peptides have been shown to display similar effects on in vitro and in vivo carbohydrate metabolism. The present study was initiated to identify and characterize the binding sites for these peptides in lung and nucleus accumbens membranes prepared from pig and guinea pig. Both tissues in either species displayed high-affinity (2-[125I]iodohistidyl10)humanCGRPα ([125I]hCGRPα) binding (IC50 = 0.4–7.7 nM), which was displaced by hCGRP8–37α with equally high affinity (IC50 = 0.4–7.3 nM). High-affinity binding for [125I]Bolton-Hunter human amylin ([125I]BH-h-amylin) was also observed in these tissues (IC50 = 0.2–6.0 nM). In membranes from the nucleus accumbens of both species, salmon calcitonin competed for amylin binding sites with high affinity (IC50 = 0.1 nM) but was poor in competing for amylin binding in lung membranes. Rat amylin8–37 competed for [125I]hCGRPα binding with higher affinity (IC50 = 5.4 nM) compared with [125I]BH-h-amylin binding (IC50 = 200 nM) in porcine nucleus accumbens, whereas in guinea pig nucleus accumbens, the IC50 values for rat amylin8–37 were 117 and 12 nM against [125I]hCGRPα and [125I]BH-h-amylin, respectively. Also, functional studies evaluating the activation of adenylate cyclase and generation of cyclic AMP in response to these agonists indicated that hCGRPα (EC50 = 0.3 nM), h-amylin (EC50 = 150 nM), and salmon calcitonin (EC50 = 1,000 nM) activated adenylate cyclase, resulting in increased cyclic AMP production in porcine lung membranes that was antagonized by hCGRP8–37α. The affinity of hCGRP8–37α was similar for all three peptides. The cyclic AMP responses to amylin and salmon calcitonin were significantly (p < 0.05) lower than that of hCGRPα and not additive, suggesting that they are acting as partial agonists at the same CGRP1-type receptor in porcine lung membranes. Similar observations were made for guinea pig lung membranes. However, human amylin and salmon calcitonin were weaker than hCGRPα in activating lung adenylate cyclase. None of these peptides activated adenylate cyclase in membranes prepared from the nucleus accumbens of both species. The data from these studies demonstrate both species and tissue differences in the existence of distinct CGRP and amylin binding sites and present a potential opportunity to study further CGRP and amylin receptor subtypes.  相似文献   

11.
Two series of peptides with hydrophobic side chains, Nps-(L -Leu-L -Leu-L -Ala)n-OEt and Nps-(L -Met-L -Met-L Leu)n-OEt (n = 1–6), were synthesized by the fragment condensation method using dicyclohexylcarbodiimide in the presence of N-hydroxysuccinimide. The tripeptide fragments were prepared stepwise by dicyclohexylcarbodiimide-mediated reaction of Nps-amino acids, which were synthesized by an improved rapid procedure.  相似文献   

12.
Lysine oligopeptides. Preparation by ion-exchange chromatography   总被引:3,自引:0,他引:3  
The preparation of L -lysine peptides (Lysn, n = 2–14) from polyL -lysine is described. Fractionation by ion-exchange column chromatography of poly-L -lysine hydrolysates on a preparative scale resulted in 0.2–1.0 g quantities of individual members of the poly-L -lysine series. The peptides isolated proved to be analytically pure and the optical configuration was fully retained, as demonstrated by complete enzymic digestion. Peptides higher than n = 14 were also prepared. They consisted of oligolysine groups of narrow and accurately determined size distribution. Potentiometric titrations were used both to characterize the products and to demonstrate the characteristic dependence of the dissociation constants on size of the peptide.  相似文献   

13.
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   

14.
The thermal triple helix–coil transition of covalently bridged collagenlike peptides with repeating sequences of (Ala-Gly-Pro)n, n = 5–15, was studied optically. The peptides were soluble in water/acetic acid (99:1) and were found to form triple-helical structures in this solvent system beginning with n = 8. The thermodynamic analysis of the transition equilibrium curves for n = 9–13 yielded the parameters ΔH°s = ?7.0 kJ per tripeptide unit, ΔS°s = ?23.1 J deg?1 mol?1 per tripeptide unit for the coil-to-helix transition, and the apparent nucleation parameter σ ? 5 × 10?2. It was suggested through double-jump temperature experiments that the rate-limiting step during refolding is not only influenced by the difficulties of nucleation, but also by cistrans isomerization of the Gly-Pro peptide bond.  相似文献   

15.
De novo design of peptides and proteins has recently surfaced as an approach for investigating protein structure and function. This approach vitally tests our knowledge of protein folding and function, while also laying the groundwork for the fabrication of proteins with properties not precedented in nature. The success relies heavily on the ability to design relatively short peptides that can espouse stable secondary structures. To this end, substitution with α,β‐didehydroamino acids, especially α,β‐didehydrophenylalanine (ΔzPhe), comes in use for spawning well‐defined structural motifs. Introduction of ΔPhe induces β‐bends in small and 310‐helices in longer peptide sequences. The present work aims to investigate the effect of nature and the number of amino acids interspersed between two ΔPhe residues in two model undecapeptides, Ac‐Gly‐Ala‐ΔPhe‐Ile‐Val‐ΔPhe‐Ile‐Val‐ΔPhe‐Ala‐Gly‐NH2 (I) and Boc‐Val‐ΔPhe‐Phe‐Ala‐Phe‐ΔPhe‐Phe‐Leu‐Ala‐ΔPhe‐Gly‐OMe (II). Peptide I was synthesized using solid‐phase chemistry and characterized using circular dichroism spectroscopy. Peptide II was synthesized using solution‐phase chemistry and characterized using circular dichroism and nuclear magnetic resonance spectroscopy. Peptide I was designed to examine the effect of incorporating β‐strand‐favoring residues like valine and isoleucine as spacers between two ΔPhe residues on the final conformation of the resulting peptide. Circular dichroism studies on this peptide have shown the existence of a 310‐helical conformation. Peptide II possesses three amino acids as spacers between ΔPhe residues and has been reported to adopt a mixed 310/α‐helical conformation using circular dichroism and nuclear magnetic resonance spectroscopy studies. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
The preparation of the co-oligopeptides of the series H-Gly-Phe-(Gly)n-Trp-Gly-OH (n = 0, 1, 2) and of other free peptides of glycine, L -tryptophan, and L -phenylalanine is reported. The syntheses have been carried out by conventional methods, using N-hydroxysuccinimide esters for the coupling steps. The ultraviolet absorption properties of the free peptides have been investigated in water. No hypo- or hyperchromicity was found for the aromatic chromophores, with the exception of H-Gly-Phe-Trp-OH, which shows a small but significant hypochromicity. The contribution of the peptide bond to the molar absorptivity in the far ultraviolet has been separated from that of the side chain plus the ? COO? group by plotting the measured molar absorptivity ? of the farthest accessible uv maximum as a function of the number of peptide bonds (nA). The peptide bond contribution proved to be independent of nA in the range nA = 1–5, thus ruling out the onset of helical conformations in the longer chain peptides.  相似文献   

17.
In order to investigate the role of each amino acid residue in determining the secondary structure of the transmembrane segment of membrane proteins in a lipid bilayer, we made a conformational analysis by CD for lipid-soluble homooligopeptides, benzyloxycarbonyl-(Z-) Aaan-OEt (n = 5-7), composed of Ala, Leu, Val, and Phe, in three media of trifluoroethanol, sodium dodecyl sulfaie micelle, and phospholipid liposomes. The lipid-peptide interaction was also studied through the observation of bilayer phase transition by differential scanning cahrimetry (DSC). The CD studies showed that peptides except for Phe oligomers are present as a mainly random structure in trifluoroethanol, as a mixture of α-helix, β-sheet, β-turn, and /or random in micelles above the critical micellization concentration and preferably as an extended structure of α-helical or β-structure in dipalmitoyl-D,L -α-phosphatidylcholine (DPPC) liposomes of gel state. That the β-structure content of Val oligomers in lipid bilayers is much higher than that in micelles and the oligopeptides of Leu (n = 7) and Ala (n = 6) can take an α-helical structure with one to two turns in lipid bilayers despite their short chain lengths indicates that lipid bilayers can stabilize the extended structure of both α-helical and β-structures of the peptides. The DSC study for bilayer phase transition of DPPC / peptide mixtures showed that the Leu oligomer virtually affects neither the temperature nor the enthalpy of the transition, while Val and Ala oligomers slightly reduce the transition enthalpy without altering the transition temperature. In contrast, the Phe oligomer affects the phase transition in much more complicated manner. The decreasing tendency of the transition enthalpy was more pronounced for the Ala oligomer as compared with the Leu and Val oligomers, which means that the isopropyl group of the side chain has a less perturbing effect on the lipid acyl chain than the methyl group of Ala. © 1995 John Wiley & Sons, Inc.  相似文献   

18.
The use of 1H-nmr spectroscopy is demonstrated to be a useful analytical method to characterize the structure of synthetic peptides attached to soluble, macromolecular polyoxyethylene (POE) supports in the liquid-phase method (LPM) of peptide synthesis. We report an extensive 360-MHz 1H-nmr study of POE-bound homo-oligo-L -methionine peptides. A combination of high field and selective saturation or Redfield pulse methods allows resolution of individual backbone NH and α-CH resonances of dilute peptides in the presence of strong resonances from macromolecular POE and/or protonated solvents. The nmr spectra for the POE-bound peptides in CDCl3 are qualitatively similar to those of the low-molecular-weight Boc-L -Metn-OMe peptide esters. This corroborates other observations that POE has little effect on peptide stucture. The backbone α-CH region of peptides is overlapped by signals from the terminal oxyethylene group of POE, but the peptide side-chain and low-field backbone NH resonances are well resolved. In trifluoroethanol the Boc-(L -Met)n-NH-POE heptamer and octamer adopt the right-handed α-helical structure, and the present nmr studies provide evidence for two strong intramolecular hydrogen bonds to stabilize the helices. In water, the N-deblocked derivatives, (L -Met)n-NH-POE oligomers adopt β-sheet structure and manifest well-resolved nonequivalent NH resonances with 6–7 Hz 3JNH-CH coupling constants.  相似文献   

19.
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   

20.
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