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 405 毫秒
1.
α,β‐Dehydroamino acid esters occur in nature. To investigate their conformational properties, a systematic theoretical analysis was performed on the model molecules Ac‐ΔXaa‐OMe [ΔXaa = ΔAla, (E)‐ΔAbu, (Z)‐ΔAbu, ΔVal] at the B3LYP/6‐311+ + G(d,p) level in the gas phase as well as in chloroform and water solutions with the self‐consistent reaction field‐polarisable continuum model method. The Fourier transform IR spectra in CCl4 and CHCl3 have been analysed as well as the analogous solid state conformations drawn from The Cambridge Structural Database. The ΔAla residue has a considerable tendency to adopt planar conformations C5 (?, ψ ≈ ? 180°, 180°) and β2 (?, ψ ≈ ? 180°, 0°), regardless of the environment. The ΔVal residue prefers the conformation β2 (?, ψ ≈ ? 120°, 0°) in a low polar environment, but the conformations α (?, ψ ≈ ? 55°, 35°) and β (?, ψ ≈ ? 55°, 145°) when the polarity increases. The ΔAbu residues reveal intermediate properties, but their conformational dispositions depend on configuration of the side chain of residue: (E)‐ΔAbu is similar to ΔAla, whereas (Z)‐ΔAbu to ΔVal. Results indicate that the low‐energy conformation β2 is the characteristic feature of dehydroamino acid esters. The studied molecules constitute conformational patterns for dehydroamino acid esters with various side chain substituents in either or both Z and E positions. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
Buczek A  Wałęsa R  Broda MA 《Biopolymers》2012,97(7):518-528
The tendency to adopt β‐turn conformation by model dipeptides with α,β‐dehydrophenylalanine (ΔPhe) residue in the gas phase and in solution is investigated by theoretical methods. We pay special attention to a dependence of conformational properties on the side‐chain configuration of dehydro residue and the influence of N‐methylation on β‐turn stability. An extensive computational study of the conformational preferences of Z and E isomers of dipeptides Ac‐Gly‐(E/Z)‐ΔPhe‐NHMe ( 1a / 1b ) and Ac‐Gly‐(E/Z)‐ΔPhe‐NMe2 ( 2a / 2b ) by B3LYP/6‐311++G(d,p) and MP2/6‐311++G(d,p) methods is reported. It is shown that, in agreement with experimental data, Ac‐Gly‐(Z)‐ΔPhe‐NHMe has a great tendency to adopt β‐turn conformation. In the gas phase the type II β‐turn is preferred, whereas in the polar environment, the type I. On the other hand, dehydro residue in Ac‐Gly‐(E)‐ΔPhe‐NHMe has a preference to adopt extended conformations in all environments. N‐methylation of C‐terminal amide group, which prevents the formation of 1←4 intramolecular hydrogen bond, change dramatically the conformational properties of studied dehydropeptides. Especially, the tendency to adopt β‐turn conformations is much weaker for the N‐methylated Z isomer (Ac‐Gly‐(Z)‐ΔPhe‐NMe2), both in vacuo and in the polar environment. On the contrary, N‐methylated E isomer (Ac‐Gly‐(E)‐ΔPhe‐NMe2) can easier adopt β‐turn conformation, but the backbone torsion angles (?1, ψ1, ?2, ψ2) are off the limits for common β‐turn types. © 2012 Wiley Periodicals, Inc. Biopolymers 97:518–528, 2012.  相似文献   

3.
Conformations of two pairs of dehydropeptides with the opposite configuration of the ΔPhe residue, Boc‐Gly‐ΔZPhe‐Gly‐Phe‐OMe ( Z‐ OMe ), Boc‐Gly‐ΔEPhe‐Gly‐Phe‐OMe ( E‐ OMe ), Boc‐Gly‐ΔZPhe‐Gly‐Phe‐p‐NA ( Z‐p‐ NA ), and Boc‐Gly‐ΔEPhe‐Gly‐Phe‐p‐NA ( E‐p‐ NA ) were compared on the basis of CD and NMR studies in MeOH, trifluoroethanol (TFE), MeCN, chloroform, and dimethylsulfoxide (DMSO). The CD results were used as the additional input data for the NMR‐based determination of the detailed solution conformations of the peptides. It was found that E‐ OMe is unordered and Z‐ OMe , Z‐p‐ NA , and E‐p‐ NA adopt the β‐turn conformation. There are two overlapping β‐turns in each of those peptides: type II and type III′ in Z‐ OMe and Z‐p‐ NA , and two type III in E‐p‐ NA . The ordered structure‐inducing properties of ΔZPhe and ΔEPhe in the peptides studied depend on the C‐terminal blocking group. In methyl esters, the ΔZPhe residue is a strong inducer of ordered conformations whereas the ΔEPhe one has no such properties. In p‐nitroanilides, both isomers of ΔPhe cause the peptides to adopt ordered structures to a similar extent. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 1055–1064, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

4.
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   

5.
Effective peptidomimetics should posses structural rigidity and appropriate interaction pattern leading to potential spatial and electronic matching to the target receptor site. Rational design of such small bioactive molecules could push chemical synthesis and molecular modeling toward faster progress in medicinal chemistry. Conformational properties of N‐t‐butoxycarbonyl‐glycine‐(E/Z)‐dehydrophenylalanine N′,N′‐dimethylamides (Boc‐Gly‐(E/Z)‐ΔPhe‐NMe2) in chloroform were studied by NMR and IR spectroscopy. The experimental findings were supported by extensive calculations at DFT(B3LYP, M06‐2X) and MP2 levels of theory and the β‐turn tendency for both isomers of the studied dipeptide were determined in vacuum and in solution. The theoretical data and experimental IR results were used as an additional information for the NMR‐based determination of the detailed solution conformations of the peptides. The obtained results reveal that N‐methylation of C‐terminal amide group changes dramatically the conformational properties of studied dehydropeptides. Theoretical conformational analysis reveals that the tendency to adopt β‐turn conformations is much weaker for the N‐methylated Z isomer (Boc‐Gly‐(Z)‐ΔPhe‐NMe2), both in vacuum and in polar environment. On the contrary, N‐methylated E isomer (Boc‐Gly‐(E)‐ΔPhe‐NMe2) can easily adopt β‐turn conformation, but the backbone torsion angles (φ1, ψ1, φ2, ψ2) are off the limits for common β‐turn types. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 28–40, 2014.  相似文献   

6.
Young Kee Kang  In Kee Yoo 《Biopolymers》2014,101(11):1077-1087
Conformational preferences of 9‐ and 14‐helix foldamers have been studied for γ‐dipeptides of 2‐aminocyclohexylacetic acid (γAc6a) residues such as Ac‐(γAc6a)2‐NHMe ( 1 ), Ac‐(Cα‐Et‐γAc6a)2‐NHMe ( 2 ), Ac‐(γAc6a)2‐NHBn ( 3 ), and Ac‐(Cα‐Et‐γAc6a)2‐NHBn ( 4 ) at the M06‐2X/cc‐pVTZ//M06‐2X/6‐31 + G(d) level of theory to explore the influence of substituents on their conformational preferences. In the gas phase, the 9‐helix foldamer H9 and 14‐helix foldamer H14‐z are found to be most preferred for dipeptides 2 and 4 , respectively, as for dipeptides 1 and 3 , which indicates no remarkable influence of the Cα‐ethyl substitution on conformational preferences. The benzyl substitution at the C‐terminal end lead H14‐z to be the most preferred conformer for dipeptides 3 and 4 , whereas it is H9 for dipeptides 1 and 2 , which can be ascribed to the favored C? H···π interactions between the cyclohexyl group of the first residue and the C‐terminal benzyl group. There are only marginal changes in backbone structures and the distances and angles of H‐bonds for all local minima by Cα‐ethyl and/or benzyl substitutions. Although vibrational frequencies and intensities of the dipeptide 4 calculated at both M06‐2X/6‐31 + G(d) and M05‐2X/6‐31 + G(d) levels of theory are consistent with observed results in the gas phase, H14‐z is predicted to be most preferred by ΔG only at the former level of theory. Hydration did not bring the significant changes in backbone structures of helix foldamers for both dipeptide 1 and 4 . It is expected that the different substitutions at the C‐terminal end lead to the different helix foldamers, which may increase the resistance of helical structures to proteolysis and provide the more surface to the helical structures suitable for molecular recognition. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 1077–1087, 2014.  相似文献   

7.
Electron spin resonance (ESR), 1H‐NMR, voltage and resistance experiments were performed to explore structural and dynamic changes of Egg Yolk Lecithin (EYL) bilayer upon addition of model peptides. Two of them are phenylalanine (Phe) derivatives, Ac‐Phe‐NHMe ( 1 ) and Ac‐Phe‐NMe2 ( 2 ), and the third one, Ac‐(Z)‐ΔPhe‐NMe2 ( 3 ), is a derivative of (Z)‐α,β‐dehydrophenylalanine. The ESR results revealed that all compounds reduced the fluidity of liposome's membrane, and the highest activity was observed for compound 2 with N‐methylated C‐terminal amide bond (Ac‐Phe‐NMe2). This compound, being the most hydrophobic, penetrates easily through biological membranes. This was also observed in voltage and resistance studies. 1H‐NMR studies provided a sound evidence on H‐bond interactions between the studied diamides and lecithin polar head. The most significant changes in H‐atom chemical shifts and spin‐lattice relaxation times T1 were observed for compound 1 . Our experimental studies were supported by theoretical calculations. Complexes EYL? Ac‐Phe‐NMe2 and EYL? Ac‐(Z)‐ΔPhe‐NMe2, stabilized by NH???O or/and CH???O H‐bonds were created and optimized at M06‐2X/6‐31G(d) level of theory in vacuo and in H2O environment. According to our molecular‐modeling studies, the most probable lecithin site of H‐bond interaction with studied diamides is the negatively charged O‐atom in phosphate group which acts as H‐atom acceptor. Moreover, the highest binding energy to hydrocarbon chains were observed in the case of Ac‐Phe‐NMe2 ( 2 ).  相似文献   

8.
The continuously growing interest in the understanding of peptide folding led to the conformational investigation of methylamides of N‐acetyl‐amino acids as diamide models. Here we report the results of detailed conformational analysis on Ac‐Pro‐NHMe and Ac‐β‐HPro‐NHMe diamides. These compounds were analyzed by experimental and computational methods, the conformational distributions obtained by Density Functional Theory (DFT) calculations for isolated and solvated diamide compounds are discussed. The conformational preference of proline‐containing diamide compounds as a function of the ambience was observed by a number of chiroptical spectroscopic techniques, such as vibrational circular dichroism (VCD), electronic circular dichroism (ECD), Raman optical activity (ROA) spectroscopy, and additionally by single crystal X‐ray diffraction analyses. Based on a comparison between Ac‐Pro‐NHMe and Ac‐β‐HPro‐NHMe, one can conclude that due to the greater conformational freedom of the β‐HPro derivative, Ac‐β‐HPro‐NHMe shows different behavior in solid‐ and solution‐phase, as well. Ac‐β‐HPro‐NHMe tends to form cis Ac‐β‐HPro amide conformation in water, dichloromethane, and acetonitrile in contrast to its α‐Pro analog. On the other hand, the crystal structure of the β‐HPro compound cannot be related to any of the conformers obtained in vacuum and solution while the X‐ray structure of Ac‐Pro‐NHMe was identified as tαL–, which is a trans Ac‐Pro amide containing conformer also predominant in polar solvents. Chirality 26:228–242, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

9.
Dehydroamino acids are non‐coded amino acids that offer unique conformational properties. Dehydrophenylalanine (ΔPhe) is most commonly used to modify bioactive peptides to constrain the topography of the phenyl ring in the side chain, which commonly serves as a pharmacophore. The Ramachandran maps (in the gas phase and in CHCl3 mimicking environments) of ΔPhe analogues with methyl groups at the β position of the side chain as well as at the C‐terminal amide were calculated using the B3LYP/6‐31 + G** method. Unexpectedly, β‐methylation alone results in an increase of conformational freedom of the affected ΔPhe residue. However, further modification by introducing an additional methyl group at C‐terminal methyl amide results in a steric crowding that fixes the torsion angle ψ of all conformers to the value 123°, regardless of the Z or E position of the phenyl ring. The number of conformers is reduced and the accessible conformational space of the residues is very limited. In particular, (Z)‐Δ(βMe)Phe with the tertiary C‐terminal amide can be classified as the amino acid derivative that has a single conformational state as it seems to adopt only the β conformation. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
Summary α,β-Dehydroamino acids are useful peptide modifiers. However, their stereoelectronic properties still remain insufficiently recognized. Based on FTIR experiments in the range ofv s(N-H), AI, AII andv s(Cα=Cβ) and ab initio calculations with B3LYP/6–31G*, we studied the solution conformational preferences and the amide electron density perturbation of Ac-ΔXaa-NHMe, where ΔXaa=ΔAla, (E)-ΔAbu, (Z)-ΔAbu, (Z)-ΔLeu, (Z)-ΔPhe and ΔVal. Each of these dehydroamides adopts a C5 structure, which in Ac-ΔAla-NHMe is fully extended and accompanied by the strong C5 hydrogen bond. Interaction with bond Cα=Cβ lessens the amidic resonance within the flanking amide groups. TheN-terminal C=O bond is noticeably shorter, both amide bonds are longer than the corresponding bonds in the saturated entities and the N-terminal amide system is distorted. Ac-ΔAla-NHMe constitutes an exception. ItsC-terminal amide bond is shorter than the standard one and both amide systems are ideally planar. Ac-(E)-ΔAbu-NHMe shares stereoelectronic features with both Ac-ΔAla-NHMe and (Z)-dehydroamides.  相似文献   

11.
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   

12.
Extracts of the female sex pheromone gland of the carpenterworm moth, Holcocerus vicarius (Walker) (Lepidoptera: Cossidae), a pest of Ulmus pumila L. (Ulmaceae), were found to contain Z7‐tetradecenyl acetate (Z7‐14Ac), E3‐tetradecenyl acetate (E3‐14Ac), (Z3,E5)‐tetradecenyl acetate (Z3,E5‐14Ac), and Z7‐tetradecenyl alcohol (Z7‐14OH) by coupled gas chromatographic‐electroantennographic detection (GC‐EAD) and coupled gas chromatography‐mass spectrometry (GC‐MS). Field trapping studies with impregnated rubber septa indicated that Z7‐14Ac was essential for attraction of males of H. vicarius. However, the most attractive blend contained Z7‐14Ac, E3‐14Ac, Z3,E5‐14Ac, and Z7‐14OH in a 50:22:17:10 ratio. Our results demonstrated that a blend of Z7‐14Ac, E3‐14Ac, Z3,E5‐14Ac, and Z7‐14OH represented the sex pheromone of H. vicarius. The optimized four‐component lure blend may be useful for monitoring H. vicarius infestations and mating disruption.  相似文献   

13.
The four possible isomers of tetradeca‐4,8‐dien‐1‐yl acetate and corresponding alcohols were synthesized stereoselectively by synthetic routes employing Wittig coupling reaction for the preparation of (Z,E)‐ and (Z,Z)‐isomers, and alkylation of terminal alkynes for the preparation of (E,E)‐ and (E,Z)‐isomers as the key steps. Synthetic products were characterized by 13C‐ and 1H‐NMR spectroscopy as well as mass‐spectrometric methods. All four isomers gave distinctive mass spectra where m/z 81 fragments clearly dominated. Elution order, followed by retention index presented in parenthesis, of tetradeca‐4,8‐dien‐1‐ols was determined as (Z,Z) (2082.1), (Z,E) (2082.8), (E,E) (2083.1), and (E,Z) (2083.2) from unpolar SPB‐1 column, and as (E,E) (2210.2), (Z,E) (2222.1), (E,Z) (2223.4), and (Z,Z) (2224.7) from polar DB‐WAX column. The isomers of tetradeca‐4,8‐dien‐1‐yl acetates eluted in the order of (Z,Z) (2176.1), (Z,E) (2178.4), (E,Z) (2185.9), and (E,E) (2186.4) from SPB‐1, and (Z,E) (2124.3), (E,E) (2157.7), (Z,Z) (2128.9), and (E,Z) (2135.9) from DB‐WAX columns. Field‐screening tests for attractiveness of tetradeca‐4,8‐dien‐1‐yl acetates revealed that (4Z,8E)‐tetradeca‐4,8‐dien‐1‐yl acetate significantly attracted Phyllonorycter coryli and Chrysoesthia drurella males. (4E,8E)‐Tetradeca‐4,8‐dien‐1‐yl acetate was the most efficient attractant for Ph. esperella and Ph. saportella males, and (4E,8Z)‐tetradeca‐4,8‐dien‐1‐yl acetate was attractive to Ph. cerasicolella males.  相似文献   

14.
N‐(tert‐butyloxycarbonyl) or N‐(9‐fluorenylmethoxycarbonyl) dipeptides with C‐terminal (Z)‐α,β‐didehydrophenylalanine (?ZPhe), (Z)‐α,β‐didehydrotyrosine (?ZTyr), (Z)‐α,β‐didehydrotryptophan (?ZTrp), (Z)‐α,β‐didehydromethionine (?ZMet), (Z)‐α,β‐didehydroleucine (?ZLeu), and (Z/E)‐α,β‐didehydroisoleucine (?Z/EIle) were synthesised from their saturated analogues via oxidation of intermediate 2,5‐disubstituted‐oxazol‐5‐(4H)‐ones (also known as azlactones) with pyridinium tribromide followed by opening of the produced unsaturated oxazol‐5‐(4H)‐one derivatives in organic‐aqueous solution with a catalytic amount of trifluoroacetic acid or by a basic hydrolysis. In all cases, a very strong preference for Z isomers of α,β‐didehydro‐α‐amino acid residues was observed except of the ΔIle, which was obtained as the equimolar mixture of Z and E isomers. Reasons for the (Z)‐stereoselectivity and the increased stability of the aromatic α,β‐didehydro‐α‐amino acid residue oxazol‐5‐(4H)‐ones over the corresponding aliphatic ones are also discussed. It is the first use of such a procedure to synthesise peptides with the C‐terminal unsaturated residues and a peptide with 2 consecutive ΔPhe residues. This approach is very effective especially in the synthesis of peptides with aliphatic α,β‐didehydro‐α‐amino acid residues that are difficult to obtain by other methods. It allowed the first synthesis of the ?Met residue. It is also more cost‐effective and less laborious than other synthesis protocols. The dipeptide building blocks obtained were used in the solid‐phase synthesis of model peptides on a polystyrene‐based solid support. Peptides containing aromatic α,β‐didehydro‐α‐amino acid residues were obtained with PyBOP or TBTU as a coupling agent with good yields and purities. In the case of aliphatic α,β‐didehydro‐α‐amino acid residues, a good efficiency was achieved only with DPPA as a coupling agent.  相似文献   

15.
16.
Conformational preferences and prolyl cis?trans isomerizations of the (2S,4S)‐4‐methylproline (4S‐MePro) and (2S,4R)‐4‐methylproline (4R‐MePro) residues are explored at the M06‐2X/cc‐pVTZ//M06‐2X/6‐31+G(d) level of theory in the gas phase and in water, where solvation free energies were calculated using the implicit SMD model. In the gas phase, the down‐puckered γ‐turn structure with the trans prolyl peptide bond is most preferred for both Ac‐4S‐MePro‐NHMe and Ac‐4R‐MePro‐NHMe, in which the C7 hydrogen bond between two terminal groups seems to play a role, as found for Ac‐Pro‐NHMe. Because of the C7 hydrogen bonds weakened by the favorable direct interactions between the backbone C?O and H? N groups and water molecules, the 4S‐MePro residue has a strong preference of the up‐puckered polyproline II (PPII) structure over the down‐puckered PPII structure in water, whereas the latter somewhat prevails over the former for the 4R‐MePro residue. However, these two structures are nearly equally populated for Ac‐Pro‐NHMe. The calculated populations for the backbone structures of Ac‐4S‐MePro‐NHMe and Ac‐4R‐MePro‐NHMe in water are reasonably consistent with CD and NMR experiments. In particular, our calculated results on the puckering preference of the 4S‐MePro and 4R‐MePro residues with the PPII structures are in accord with the observed results for the stability of the (X‐Y‐Gly)7 triple helix with X = 4R‐MePro or Pro and Y = 4S‐MePro or Pro. The calculated rotational barriers indicate that the cis?trans isomerization may in common proceed through the anticlockwise rotation for Ac‐4S‐MePro‐NHMe, Ac‐4R‐MePro‐NHMe, and Ac‐Pro‐NHMe in water. The lowest rotational barriers become higher by 0.24?1.43 kcal/mol for Ac‐4S‐MePro‐NHMe and Ac‐4R‐MePro‐NHMe than those for Ac‐Pro‐NHMe in water. © 2010 Wiley Periodicals, Inc. Biopolymers 95: 51–61, 2011.  相似文献   

17.
The influence of aqueous environment on the main‐chain conformation (ω0, ?, and ψ dihedral angles) of two model peptoids: N‐acetyl‐N‐methylglycine N’‐methylamide (Ac‐N(Me)‐Gly‐NHMe) ( 1 ) and N‐acetyl‐N‐methylglycine N’,N’‐dimethylamide (Ac‐N(Me)‐Gly‐NMe2) ( 2 ) was investigated by MP2/6‐311++G(d,p) method. The Ramachandran maps of both studied molecules with cis and trans configuration of the N‐terminal amide bond in the gas phase and in water environment were obtained and all energy minima localized. The polarizable continuum model was applied to estimate the solvation effect on conformation. Energy minima of the Ac‐N(Me)‐Gly‐NHMe and Ac‐N(Me)‐Gly‐NMe2 have been analyzed in terms of the possible hydrogen bonds and C = O dipole attraction. To validate the theoretical results obtained, conformations of the similar structures gathered in the Cambridge Crystallographic Data Centre were analyzed. Obtained results indicate that aqueous environment in model peptoids 1 and 2 favors the conformation F (? and ψ = ?70º, 180º), and additionally significantly increases the percentage of structures with cis configuration of N‐terminal amide bond in studied compounds. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
Female European corn borer, Ostrinia nubilalis, produce and males respond to sex pheromone blends with either E‐ or Z‐Δ11‐tetradecenyl acetate as the major component. E‐ and Z‐race populations are sympatric in the Eastern United States, Southeastern Canada, and the Mediterranean region of Europe. The E‐ and Z‐pheromone races of O. nubilalis are models for incipient species formation, but hybridization frequencies within natural populations remain obscure due to lack of a high‐throughput phenotyping method. Lassance et al. previously identified a pheromone gland‐expressed fatty‐acyl reductase gene (pgfar) that controls the ratio of Δ11‐tetradecenyl acetate stereoisomers. We identified three single nucleotide polymorphism (SNP) markers within pgfar that are differentially fixed between E‐ and Z‐race females, and that are ≥98.2% correlated with female pheromone ratios measured by gas chromatography. Genotypic data from locations in the United States demonstrated that pgfar‐z alleles were fixed within historically allopatric Z‐pheromone race populations in the Midwest, and that hybrid frequency ranged from 0.00 to 0.42 within 11 sympatric sites where the two races co‐occur in the Eastern United States (mean hybridization frequency or heterozygosity (HO) = 0.226 ± 0.279). Estimates of hybridization between the E‐ and Z‐races are important for understanding the dynamics involved in maintaining race integrity, and are consistent with previous estimates of low levels of genetic divergence between E‐ and Z‐races and the presence of weak prezygotic mating barriers.  相似文献   

19.
A set of cyclic tetrapeptides of the general form cyclo (Boc‐Cys‐Pro‐ X ‐Cys‐OMe) with X being L‐ / D‐Ala , L‐ / D‐Val , and L‐ / D‐Trp was synthesized. These peptides serve as model systems for structure elucidation in solution and feature a variety of structural motifs — namely a β‐turn with intramolecular hydrogen bonding interactions, cis/trans isomerism, and a disulphide bond. In this work, we performed a comprehensive structural analysis focussing on their β‐turn conformational preferences using NMR, VCD, and Raman spectroscopy. Our results provide evidence for a strong influence of a single stereocenter on the structures of the peptides whereas solvent polarity does not significantly affect them. Additionally, the solid state conformational preferences were studied by crystal structure analysis. Overall, a general trend for the conformational preferences of this set of peptides can be concluded from the results of the complementary investigations.  相似文献   

20.
The major sex pheromone compound of the spotted tentiform leafminer, Phyllonorycter blancardella (F.) (Lepidoptera: Gracillariidae), from Ontario, Canada, was identified as (10E)‐dodecen‐1‐yl acetate (E10‐12:Ac) using chemical analysis and field trapping experiments. The minor compounds (10E)‐dodecen‐1‐ol (E10‐12:OH) (4.6%), dodecan‐1‐ol (12:OH) (2.3%), and (10Z)‐dodecen‐1‐yl acetate (Z10‐12:Ac) 1.6% were also identified. The dienic acetate (4E,10E)‐dodecadien‐1‐yl acetate (E4,E10‐12:Ac), a compound reported to be attractive to P. blancardella, was not found in the glands of this population. A two‐component blend of the major and one of each the three minor compounds, in ratios similar to those found in the sex pheromone gland, did not increase the attractiveness of traps baited with synthetic pheromone. The minor compounds E10‐12:OH and 12:OH were not attractive to P. blancardella when tested individually. Z10‐12:Ac was attractive to P. blancardella, although traps baited with this compound captured only 2% of the moths that were captured in traps baited with the main compound. A four‐component blend of the major and each of the three minor compounds (100 : 1 : 1 : 1) was not more attractive than the major compound alone. The related species Phyllonorycter mespilella was captured in traps baited with E10‐12:Ac.  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