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.
The conformations of the phytotoxic cyclic tetrapeptide tentoxin [cyclo-(L -MeAla1-L -Leu2-MePhe[(Z)Δ]3-Gly4 )] have been studied in aqueous solution by two-dimensional proton nmr at various temperatures. Contrary to what is observed in chloroform, tentoxin exhibits multiple exchanging conformations in water. Aggregation phenomena were also observed. Four conformations with different proportions (51, 37, 8, and 4%) were observed at ?5°C. Models were constructed from nmr parameters and restrained molecular dynamics simulations. All the models exhibit cis-trans-cis-trans conformation of the amide bond sequence. The conversion from one form to another is accomplished by a conformational peptide flip consisting of a 180° rotation of a nonmethylated peptide bond. © 1995 John Wiley & Sons, Inc.  相似文献   

2.
220-MHz NMR spectra of various poly (N-alkylamino acid)s are investigated. Spectra of polysarcosine recorded in various solvents showed fine splittings of the methyl and methylene bands. Comparing the spectrum with that of its model compound, the fine structure of the methyl band of polysarcosine was assigned to four dyad sequences of the cistrans isomeric state of the main chain amide bonds. Also the methylene band was roughly divided into cis and trans bands. From the temperature dependence of the spectra of polysarcosine, a double coalescence phenomenon was observed, in which the four dyad peaks coalesced into two peaks corresponding to cis and trans, then the two peaks coalesced into one peak. Further, the approximate value of the free energy for the internal rotation of the main chain amide bond was estimated. NMR spectra of various poly(N-alkylglycine)s in methylene chloride solution were also obtained. From the comparsion of their methylene bands, the introduction of the bulky N-alkyl groups was found to increase the cis content of the amide bond.  相似文献   

3.
P R Andrews 《Biopolymers》1971,10(11):2253-2267
The molecular orbital method PCILO is applied to eight. N-monsubstituted amides. Experimentally known geometric properties are reasonably predicted by minimization of total energy with respect to molecular geometry. The same procedure shows that molecular deformations during rotation around the peptide bond significantly lower calculated barriers. Experimental heats of activation and the free-energy changes associated with cis–trans isomerism are in good agreement with those calculated, which include qualitative estimates of configurational entropy contributions to the isomerism energies. Both the calculations and revised infrared data indicate that N-phenylurethane, which has been used as a model for the cis peptide bond, should be predominantly trans. However the variations in rotational barriers and cis–trans isomerism energies among the N-monosubstituted amides provide no reason to suppose that the cis peptide bond should be excluded from stable protein conformations.  相似文献   

4.
5.
A microphotometric technique that displays rapid length changes of Spirostomum has been used to follow the variation with temperature of three kinetic parameters of myonemal contraction: contraction rate, relaxation rate and stimulus duration at threshold. In each case the exponential form of the relationship indicated that the gross rate constant might be equated with the limiting rate constant, k, of a driving chemical reaction, and from standard expressions of chemical kinetics the change in activation free energy appropriate to this reaction has been computed. The thermal dependence of contraction is described by an activation enthalpy (ΔLH?) of 21.7 kcal mol?1, and the activation entropy (ΔLS?) of 26.8 e.u. is consistent with a model of contraction requiring neutralization of fixed myonemal charges by divalent cations. The analysis of thermal dependence of relaxation gives a negative activation entropy, a result predicted for a rate-limiting reaction involving dissociation of a neutral molecule. On the other hand, values of ΔLS? and ΔLH? for relaxation fall close to an isokinetic correlation drawn in the literature from analysis of the thermal dependence of ciliary beat frequency in different organisms, and for which breakdown of an ATP-ATPase complex could be the common rate-limiting reaction. ΔLS? for stimulus duration suggests that the rate-limiting step in excitation-contraction coupling is a reaction between ions of like charge, or ion pair formation from a neutral molecule.  相似文献   

6.
Poly-L -histidine (PLH) films at different degrees of protonation were produced mid subjected to infrared spectroscopic investigation (range 4000-650 cm?1). In addition, the N-deuterated film spectra were plotted. The amide II and III bands show that the peptide group is present in the trans form. The amide I and II bands show that at 0% and 50% protonation the PLH occurs as an α-helix and at 100% protonation as a random coil with some ranges in β structure. At 0% and 50% protonation, no hydration water is bound to the backbone. At 0% protonation all NH groups are linked to each other or to water molecules via hydrogen bonds. At 50% protonation NH+?N bonds form between the imidazole rings. These protons are present in continuous energy level distribution. Such bonds with tunneling protons are extremely polarizable and between these bonds may act proton dispersion forces. The Cl? ions are bonded to the NH groups of the imidazole groups. The hydration water is bonded to the Cl?? ions and to the NH groups. At 100% protonation, hydration water is bonded also to the CO groups of the backbone. The NH groups of the backbone, like those of the rings, endeavor especially in the dry state to bond to the Cl? ions. This leads to a strong steric constraint of the random coil.  相似文献   

7.
The physicochemical parameters affecting protein unfolding in relation to peptide bond rotations are briefly reviewed. As a suitable model for the study of solvent and solute effects on amide rotation and inside-outside conformations, the 2,2′-biphenyl analog of N-benzoyl-l-phenylalanine methyl ester (I) was synthesized and resolved enzymatically with α-chymotrypsin. The optically pure substance exists as conformer Ia (R,S configuration) with an axial methoxycarbonyl in the crystalline state. Rotation about the biphenyl axis leads to the equatorial conformer Ib (S,S configuration) in various solvents. In polar solvents, rotation about the amide is rate limiting. Accurate measurements of this rotation were accomplished by following the rate of change in the maximum amplitude of the biphenyl Cotton band at 256 nm. The high sensitivity of the method allowed rate and equilibrium measurements at 10?3M in the absence of intermolecular association. Small differences of the order of 100 cal/mole in ΔG or ΔGeq could thus be detected accurately. It was found that kobs or k1 (forward step) for equilibration was linearly related (correlation coefficient of 0.96 for kobs) with ET, the solvent polarity index on Reichardt and Dimroth's scale. Rotation was slowest in water and fastest in carbon tetrachloride, δΔG, being 2.4 kcal/mole. Chaotropic anions, cations, and guanidinium chloride accelerated the rate in water. However, the inside-outside (axial-equatorial; IaIb) ratio at equilibrium did not correlate in any simple manner with the solvent ET values. Rather, correlation within groups of solvents appeared to exist. It was suggested that solvent association with the amide differs quantitatively in the inside and outside conformations. The position of the equilibrium in water was affected by chaotropic ions but not by urea or quanidinium chloride. Some possible mechanisms are briefly outlined.  相似文献   

8.
Bark of Virola sebifera used in the preparation of hallucinogenic snuffs and drinks in Venezuela has yielded N-methyl-N-formyltryptamine and N-methyl-N-acetyltryptamine, which exist in two rotameric forms reflecting hindered rotation around the carbon—nitrogen bond of the amide function. They were detected by HPLC as well as NMR. 2-Methyl-1,2,3,4-tetrahydro-β-carboline, N,N-dimethyltryptamine, its oxide and N-monomethyltryptamine were also identified.  相似文献   

9.
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   

10.
The self-association of N-acetylglycine N,N-dimethylamide, N-acetyl-L -valine N,N-dimethylamide, and N-acetyl-L -phenylalanine N,N-dimethylamide in carbon tetrachloride was investigated by using ir and 1H-nmr methods. It was concluded from ir measurements that the associated species is the dimer formed as a result of the simultaneous formation of two intermolecular hydrogen bonds. This is supported by the results of 1H-nmr measurements. Thermodynamic quantities for the association were determined from the temperature and concentration dependence of the NH proton chemical shifts of the sample solutions. Compared with the Gly derivative, L -Val and L -Phe derivatives have larger values of ?ΔH for association, which shows good correlation with ΔvNH values, the difference between the maxima of the monomer and dimer bands, obtained from ir spectra. This is due to the less stable monomer conformation and to the stronger intermolecular hydrogen bonding of the dimers in L -Val and L -Phe derivatives. The line shapes of both methyl proton resonances of L -Val residue and methylene proton resonances of L -Phe residue were found to vary with concentration and temperature of the sample solutions. These data indicate that the rotation about the Cα—Cβ bond is restricted by the steric hindrance present in the associated dimers. All these experimental results can be related to the fact that L -Val and L -Phe derivatives have a warped framework because of the bulky side chains, whereas the Gly derivative has a planar framework.  相似文献   

11.
Abstract

The design rationale for a new series of tripeptide derived angiotensin converting enzyme (ACE) inhibitors, which we term “ketomethylureas”, is described. Analogs of tripeptide substrates (i.e. N-benzoyl-Phe-Ala-Pro) in which the nitrogen atom of the scissile amide bond and the adjacent asymmetric carbon atom of the penultimate amino acid residue are formally transposed give rise to this novel class of inhibitors. The most potent ketomethylureas inhibit ACE wtih I50 values in the nM range.  相似文献   

12.
Synthesis of a peptide bond is suggested to be enzymatically catalysed in a biphasic system “water–water-immiscible organic solvent”. The pH dependence of the apparent equilibrium constant is studied for synthesis of N-acetyl-L -tryptophanyl-L -leucine amide from N-Acet-Trp andL -Leu-NH2. The reaction was performed in the biphasic system ethyl acetate plus water [from 2 to 2% (v/v)] in the presence of α-chymotrypsin. The suggested approach is preparative value: with the stoichiometric ratio of the reagents, [N-Acet-L -Trp] = [L -Leu-NH2] = 2 × 10?3M, the yield is practically 100% (in water, with other conditions being the same, the yield is not over 01.%).  相似文献   

13.
The 4,5-dimethoxy-2-mercaptobenzyl (Dmmb) group attached to a main chain amide in a peptide is easily transformed into an S-peptide via an intramolecular NS acyl shift reaction under acidic conditions, and the S-peptide produces a peptide thioester through an intermolecular thiol–thioester exchange reaction. In order to develop a method for efficiently preparing peptide thioesters based on the NS acyl shift reaction, the factors involved in this process were analyzed in detail. The general features of the transformation at the Dmmb group attached amide bond in a trifluoroacetic acid (TFA) solution and the generation of a peptide thioester were examined by 13C-NMR spectral measurements, reversed-phase (RP) HPLC analyses, mass measurements, and amino acid analyses. The methoxy group of the Dmmb group was not essential for the NS acyl shift reaction, but played a role in stabilizing the thioester form. The addition of water to the TFA solution accelerated the NS acyl shift reaction mediated by the Dmmb group and also suppressed the acid-catalyzed cleavage of the Dmmb group. A peptide thioester was produced from the S-peptide via an intermolecular thiol–thioester exchange reaction with minimal epimerization of the amino acid residue that constituted the thioester bond. Undesirable side reactions, such as the hydrolysis of the thioester bond and an S–N acyl shift reaction occurred during the synthetic process, which is a subject of further investigation.  相似文献   

14.
Conformational control in peptoids, N-substituted glycines, is crucial for the design and synthesis of biologically-active compounds and atomically-defined nanomaterials. While there are a growing number of structural studies in solution, most have been performed with conformationally-constrained short sequences (e.g., sterically-hindered sidechains or macrocyclization). Thus, the inherent degree of heterogeneity of unconstrained peptoids in solution remains largely unstudied. Here, we explored the folding landscape of a series of simple peptoid tetramers in aqueous solution by NMR spectroscopy. By incorporating specific 13C-probes into the backbone using bromoacetic acid-2-13C as a submonomer, we developed a new technique for sequential backbone assignment of peptoids based on the 1,n-Adequate pulse sequence. Unexpectedly, two of the tetramers, containing an N-(2-aminoethyl)glycine residue (Nae), had preferred conformations. NMR and molecular dynamics studies on one of the tetramers showed that the preferred conformer (52%) had a trans-cis-trans configuration about the three amide bonds. Moreover, >80% of the ensemble contained a cis amide bond at the central amide. The backbone dihedral angles observed fall directly within the expected minima in the peptoid Ramachandran plot. Analysis of this compound against similar peptoid analogs suggests that the commonly used Nae monomer plays a key role in the stabilization of peptoid structure via a side-chain-to-main-chain interaction. This discovery may offer a simple, synthetically high-yielding approach to control peptoid structure, and suggests that peptoids have strong intrinsic conformational preferences in solution. These findings should facilitate the predictive design of folded peptoid structures, and accelerate application in areas ranging from drug discovery to biomimetic nanoscience.  相似文献   

15.
The molar Kerr constants mK, molar refractions mR, and dipole moments μ are reported for the N-methylacetamides CX3CONHCH3 (X = H, CH3, F. CI, Br) and acetamides CX3CONH2 (X = H, F, Cl, Br). The components of the polarizability tensor α are deduced for N-methylacetamide and acetamide on the basis of the bond additivity approximation. This α is found to be considerably more anisotropic than was indicated in previous determinations by other methods. The data for N-methylacetamide were used to calculate mK, μ, and γ2 (anisotropy squared) of N-acetyl-N′-methylglycine amide and N-acetyl-N′-methyl-alanine amide as functions of the torsional angles (?,Ψ). The statistical mechanical averages of mK, μ, and γ2 were calculated from conformational energies obtained by the methods of Scheraga.  相似文献   

16.
Endothia parasitica protease hydrolyzes l-leucyl-l-leucine amide and l-leucyl-l-phenylalanine amide at the peptide bond. l-Phenylalanyl-l-leticine amide, N-carbobenzoxy-l-leucyl-l-phenylalanine amide, N-carbobenzoxy-l-leucyl-l-pheml-alanine, N-carbobenzoxy-l-phenylalanyl-l-valine amide, and l-leucyl-β-naphthyl-amide are not hydrolyzed. In contrast to the kinetics of hydrolysis of casein and oxidized B-chain of insulin and activation of trypsinogen by Endothia parasitica protease which are normal, reaction progress curves for hydrolysis of l-leucyl-l-leucine amide and l-leucyl-l-phenylalanine amide are sigrnoidal. Initially, the reaction rates were of the order of 0.5–2.5% of the maximum rates eventually attained. With increasing time of incubation the reaction rates became faster and faster until maximum rates were achieved. This abnormal behavior was not eliminated by recrystallization of substrate or by incubation of enzyme alone or with products of the reaction prior to addition of substrate. Addition of a new aliquot of substrate, vizl-leucyl-l-leucine amide, to the reaction prior to complete hydrolysis of all of a previous aliquot of the same substrate, or reactions containing a mixture of oxidized B chain of insulin and l-leucyl-l-leucine amide, gave normal reaction progress curves. The duration of abnormal behavior before a maximum rate was attained was a function of enzyme concentration and temperature but not of substrate concentration even though substrate was in less than saturating amounts. The reaction data follow second-order autocatalytic kinetics with respect to enzyme concentration. It is proposed that most of the enzyme is in an inactive form in absence of substrate but is rapidly converted to the active form on combination with a good substrate such as trypsinogen, casein, or oxidized B chain of insulin. However, with a poor substrate such as l-leucyl-l-leucine amide, conversion to active enzyme is mediated through formation of an active enzyme-inactive enzyme complex followed by combination with substrate and hydrolysis.  相似文献   

17.
K H Gardner  J Blackwell 《Biopolymers》1975,14(8):1581-1595
The structure of β-chitin has been refined by rigid-body least-squares methods, based on the intensity data for highly crystalline specimens from the pogonophore Oligobrachia ivanovi. The structure consists of an array of poly-N-acetyl-D -glucosamine chains all having the same sense, which are linked together in sheets by N? H … O?C hydrogen bonding of the amide groups. In addition to the O-3′? H … O-5 intramolecular hydrogen bond, analogous to that in cellulose, the CH2OH side chain forms an intrasheet hydrogen bond to the carbonyl oxygen on the next chain. This structure shows considerably better agreement between observed and calculated intensities than that possessing an intersheet hydrogen bond, as had been proposed previously. The structure is consistent with the swelling properties of β-chitin and can also be seen to be analogous to that of native cellulose.  相似文献   

18.

Mannosylphosphorylated glycans are found only in fungi, including yeast, and the elimination of mannosylphosphates from glycans is a prerequisite for yeast glyco-engineering to produce human-compatible glycoproteins. In Saccharomyces cerevisiae, MNN4 and MNN6 genes are known to play roles in mannosylphosphorylation, but disruption of these genes does not completely remove the mannosylphosphates in N-glycans. This study was performed to find unknown key gene(s) involved in N-glycan mannosylphosphorylation in S. cerevisiae. For this purpose, each of one MNN4 and five MNN6 homologous genes were deleted from the och1Δmnn1Δmnn4Δmnn6Δ strain, which lacks yeast-specific hyper-mannosylation and the immunogenic α(1,3)-mannose structure. N-glycan profile analysis of cell wall mannoproteins and a secretory recombinant protein produced in mutants showed that the MNN14 gene, an MNN4 paralog with unknown function, is essential for N-glycan mannosylphosphorylation. Double disruption of MNN4 and MNN14 genes was enough to eliminate N-glycan mannosylphosphorylation. Our results suggest that the S. cerevisiae och1Δmnn1Δmnn4Δmnn14Δ strain, in which all yeast-specific N-glycan structures including mannosylphosphorylation are abolished, may have promise as a useful platform for glyco-engineering to produce therapeutic glycoproteins with human-compatible N-glycans.

  相似文献   

19.
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   

20.
Amphiphiles with one or two amide linkages have been prepared by the reaction (A) of D-gluconic acid lactone with aliphatic amines (C6-C10) and (C) of N′-gluconoyl-ethylenediamine with alkanoic acids (C6)-(C10). Gel formation was found to occur on cooling the aqueous solutions at concentrations as low as 1–2%. Electron microscopy revealed that the gels of type A are composed of highly ordered ropes with right-handed twist, especially well developed with N-octylgluconamide. Type c substances with two amide linkages of opposite direction form gels consisting of smooth ribbons devoid of twisting. N-Methylation of the amide bond (type B and D substances) leads to a considerable increase in solubility. Gels are only formed from samples containing decanoic acid. These gels also consist of right-handed fibrillar ropes, only partially ordered with one N-methylated amide linkage (B), regularly aligned side-by-side with one N-methylated and one non-methylated amide bond (D). Gel formation and the typical morphology of the gels are discussed as arising mainly from strong intermolecular hydrogen bonds between amide linkages holding the molecules together and the influence of chiral centers of the carbohydrate chain which might be responsible for helical aggregates to be formed.  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