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 479 毫秒
1.
Twelve peptides, 1 – 12 , have been synthesized, which consist of alternating sequences of α‐ and β‐amino acid residues carrying either proteinogenic side chains or geminal dimethyl groups (Aib). Two peptides, 13 and 14 , containing 2‐methyl‐3‐aminobutanoic acid residues or a ‘random mix’ of α‐, β2‐, and β3‐amino acid moieties were also prepared. The new compounds were fully characterized by CD (Figs. 1 and 2), and 1H‐ and 13C‐NMR spectroscopy, and high‐resolution mass spectrometry (HR‐MS). In two cases, 3 and 14 , we discovered novel types of turn structures with nine‐ and ten‐membered H‐bonded rings forming the actual turns. In two other cases, 8 and 11 , we found 14/15‐helices, which had been previously disclosed in mixed α/β‐peptides containing unusual β‐amino acids with non‐proteinogenic side chains. The helices are formed by peptides containing the amino acid moiety Aib in every other position, and their backbones are primarily not held together by H‐bonds, but by the intrinsic conformations of the containing amino acid building blocks. The structures offer new possibilities of mimicking peptide–protein and protein–protein interactions (PPI).  相似文献   

2.
The S30 extract from E. coli BL21 Star (DE3) used for cell‐free protein synthesis removes a wide range of α‐amino acid protecting groups by cleaving α‐carboxyl hydrazides; methyl, benzyl, tert‐butyl, and adamantyl esters; tert‐butyl and adamantyl carboxamides; α‐amino form‐, acet‐, trifluoroacet‐, and benzamides; and side‐chain hydrazides and esters. The free amino acids are produced and incorporated into a protein under standard conditions. This approach allows the deprotection of amino acids to be carried out in situ to avoid separate processing steps. The advantages of this approach are demonstrated by the efficient incorporation of the chemically intractable (S)‐4‐fluoroleucine, (S)‐4,5‐dehydroleucine, and (2S,3R)‐4‐chlorovaline into a protein through the direct use of their respective precursors, namely, (S)‐4‐fluoroleucine hydrazide, (S)‐4,5‐dehydroleucine hydrazide, and (2S,3R)‐4‐chlorovaline methyl ester. These results also show that the fluoro‐ and dehydroleucine and the chlorovaline are incorporated into a protein by the normal biosynthetic machinery as substitutes for leucine and isoleucine, respectively.  相似文献   

3.
Molecular simulation by using force field parameters has been widely applied in the fields of peptide and protein research for various purposes. We recently proposed a new all‐atom protein force field, called the SAAP force field, which utilizes single amino acid potentials (SAAPs) as the fundamental elements. In this article, whole sets of the SAAP force field parameters in vacuo, in ether, and in water have been developed by ab initio calculation for all 20 proteinogenic amino acids and applied to Monte Carlo molecular simulation for two short peptides. The side‐chain separation approximation method was employed to obtain the SAAP parameters for the amino acids with a long side chain. Monte Carlo simulation for Met‐enkephalin (CHO‐Tyr‐Gly‐Gly‐Phe‐Met‐NH2) by using the SAAP force field revealed that the conformation in vacuo is mainly controlled by strong electrostatic interactions between the amino acid residues, while the SAAPs and the interamino acid Lennard‐Jones potentials are predominant in water. In ether, the conformation would be determined by the combination of the three components. On the other hand, the SAAP simulation for chignolin (H‐Gly‐Tyr‐Asp‐Pro‐Glu‐Thr‐Gly‐Thr‐Trp‐Gly‐OH) reasonably reproduced a native‐like β‐hairpin structure in water although the C‐terminal and side‐chain conformations were different from the native ones. It was suggested that the SAAP force field is a useful tool for analyzing conformations of polypeptides in terms of intrinsic conformational propensities of the single amino acid units. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2009  相似文献   

4.
A promising strategy for mediating protein–protein interactions is the use of non‐peptidic mimics of secondary structural protein elements, such as the α‐helix. Recent work has expanded the scope of this approach by providing proof‐of‐principle scaffolds that are conformationally biased to mimic the projection of side‐chains from one face of another common secondary structural element—the β‐strand. Herein, we present a synthetic route that has key advantages over previous work: monomers bearing an amino acid side‐chain were pre‐formed before rapid assembly to peptidomimetics through a modular, iterative strategy. The resultant oligomers of alternating pyridyl and six‐membered cyclic ureas accurately reproduce a recognition domain of several amino acid residues of a β‐strand, demonstrated herein by mimicry of the i, i+2, i+4 and i+6 residues.  相似文献   

5.
A library of 35 cyclic peptidomimetics has been prepared, by a combination of solid‐ and solution‐phase methods, which, together, scan amino acid residues 444–489 in a protruding loop in the subdomain‐III of the apical membrane antigen‐I (AMA‐I), an integral membrane protein found on the surface of Plasmodium falciparum merozoites. The mimetics each contain twelve residues from AMA‐I linked through the N‐ and C‐termini to a β‐hairpin‐inducing template, comprising the dipeptide D ‐prolineL ‐Apro (Apro=cis‐4‐amino‐L ‐proline). These peptidomimetics were each coupled via the 4‐amino group to a phospholipid, which allows their incorporation into reconstituted influenza virus‐like particles (virosomes) for immunization of mice, as well as their use in ELISA to characterize epitopes recognized by anti‐AMA‐I growth‐inhibitory monoclonal antibodies.  相似文献   

6.
The mimicry of protein‐sized β‐sheet structures with unnatural peptidic sequences (foldamers) is a considerable challenge. In this work, the de novo designed betabellin‐14 β‐sheet has been used as a template, and α→β residue mutations were carried out in the hydrophobic core (positions 12 and 19). β‐Residues with diverse structural properties were utilized: Homologous β3‐amino acids, (1R,2S)‐2‐aminocyclopentanecarboxylic acid (ACPC), (1R,2S)‐2‐aminocyclohexanecarboxylic acid (ACHC), (1R,2S)‐2‐aminocyclohex‐3‐enecarboxylic acid (ACEC), and (1S,2S,3R,5S)‐2‐amino‐6,6‐dimethylbicyclo[3.1.1]heptane‐3‐carboxylic acid (ABHC). Six α/β‐peptidic chains were constructed in both monomeric and disulfide‐linked dimeric forms. Structural studies based on circular dichroism spectroscopy, the analysis of NMR chemical shifts, and molecular dynamics simulations revealed that dimerization induced β‐sheet formation in the 64‐residue foldameric systems. Core replacement with (1R,2S)‐ACHC was found to be unique among the β‐amino acid building blocks studied because it was simultaneously able to maintain the interstrand hydrogen‐bonding network and to fit sterically into the hydrophobic interior of the β‐sandwich. The novel β‐sandwich model containing 25 % unnatural building blocks afforded protein‐like thermal denaturation behavior.  相似文献   

7.
The synthesis of bolaamphiphiles from unusual β‐amino acids or an alcohol and C12 or C20 spacers is described. Unusual β‐amino acids such as a sugar amino acid, an AZT‐derived amino acid, a norbornene amino acid, and an AZT‐derived amino alcohol were coupled with spacers under standard conditions to get the novel bolaamphiphiles 5 – 8 (Scheme 1), 12 and 13 (Scheme 2), and 17 and 20 (Scheme 3). Some of these compounds, on precipitation from MeOH/H2O, self‐assembled into organized molecular structures.  相似文献   

8.
We have shown for the first time that a natural protein (human insulin) can be acylated at the N‐terminus with a β‐amino acid (H‐β3hAla‐), in a process catalyzed by the β‐peptidyl aminopeptidase 3‐2W4‐BapA. This selective modification, which could also be applied for protein labeling and tagging, should be generally useful, also to protect peptides and proteins from attack by common aminopeptidases.  相似文献   

9.
A β‐barrel protein hybrid catalyst was prepared by covalently anchoring a Grubbs–Hoveyda type olefin metathesis catalyst at a single accessible cysteine amino acid in the barrel interior of a variant of β‐barrel transmembrane protein ferric hydroxamate uptake protein component A (FhuA). Activity of this hybrid catalyst type was demonstrated by ring‐opening metathesis polymerization of a 7‐oxanorbornene derivative in aqueous solution.  相似文献   

10.
We have changed the amino acid set of the genetic code of Escherichia coli by evolving cultures capable of growing on the synthetic noncanonical amino acid L ‐β‐(thieno[3,2‐b]pyrrolyl)alanine ([3,2]Tpa) as a sole surrogate for the canonical amino acid L ‐tryptophan (Trp). A long‐term cultivation experiment in defined synthetic media resulted in the evolution of cells capable of surviving Trp→[3,2]Tpa substitutions in their proteomes in response to the 20 899 TGG codons of the E. coli W3110 genome. These evolved bacteria with new‐to‐nature amino acid composition showed robust growth in the complete absence of Trp. Our experimental results illustrate an approach for the evolution of synthetic cells with alternative biochemical building blocks.  相似文献   

11.
The kinetics of the permanganic oxidation process of some straight chain amino acids in moderately concentrated sulfuric acid medium have been investigated using a spectrophotometric technique. Conclusive evidences have proven autocatalytic activity of Mn(II) for these reactions. It is determined that even and odd effects of the number carbon atom in a carbon chain are annihilated when it's the number of carbon atoms is increased more than of three in a noncatalytic oxidation pathway. Thus, rate constants belonging to glycine, l ‐α‐amino‐n‐butyric acid, l ‐norleucine, and l ‐α‐amino‐n‐heptanoic acid satisfy Taft's equation involving the induction factor in the noncatalytic pathway, whereas l ‐α‐amino‐n‐heptanoic acid has an odd number of carbon atom in its chain carbon. On the other hand, in the catalytic pathway, rate constants satisfy Taft' equation including inductive and steric factors, when rate constants belonging to amino acids with an even number of carbon atoms are separated from those with an odd number of carbon atoms. The oxidation process of amino acids in the noncatalytic pathway and those with the even number of carbon atoms in the carbon chain in the catalytic pathway speeds up by an increase in the length of chain that is accompanied with an increase in the carbon chain's electron‐donating characteristic. On the other hand, an increase in the length of the carbon chain is accompanied with more steric hindrance, which counteracts its electron‐donating character, thereby decreasing reaction rate in the catalytic pathway. Finally, amino acid–Mn(II) complexes were studied using a density functional theory method. Results obtained show that such a complex is less stable than reactants, namely it is formed in an endothermic reaction. The number and strength of hydrogen bonding belonging to amino acid is more than those of the amino acid–Mn(II) complex. Besides, it has been illustrated that natural bond orbital analysis and molecular orbital calculations satisfy the findings.  相似文献   

12.
An efficient four‐step synthetic strategy for cis‐2,5‐disubstituted chiral piperazines derived from amino‐acid‐based aziridines is described. The key steps in this strategy are the highly regioselective boron trifluoride diethyl etherate (BF3 ⋅ OEt2)‐mediated ring‐opening of less‐reactive N‐Ts chiral aziridines by α‐amino acid methyl ester hydrochloride followed by Mitsunobu cyclization. This protocol has been used in an attempt to construct the piperazine core framework of natural product (+)‐piperazinomycin.  相似文献   

13.
The impressive rate accelerations that enzymes display in nature often result from boosting the inherent catalytic activities of side chains by their precise positioning inside a protein binding pocket. Such fine‐tuning is also possible for catalytic unnatural amino acids. Specifically, the directed evolution of a recently described designer enzyme, which utilizes an aniline side chain to promote a model hydrazone formation reaction, is reported. Consecutive rounds of directed evolution identified several mutations in the promiscuous binding pocket, in which the unnatural amino acid is embedded in the starting catalyst. When combined, these mutations boost the turnover frequency (kcat) of the designer enzyme by almost 100‐fold. This results from strengthening the catalytic contribution of the unnatural amino acid, as the engineered designer enzymes outperform variants, in which the aniline side chain is replaced with a catalytically inactive tyrosine residue, by more than 200‐fold.  相似文献   

14.
Albumin is an important plasma antioxidant protein, contributing to protecting mechanisms of cellular and regulatory long‐lived proteins. The metal‐catalyzed oxidation (MCO) of proteins plays an important role during oxidative stress. In this study, we examine the oxidative modification of albumin using an MCO in vitro system. Mass spectrometry, combined with off‐line nano‐liquid chromatography, was used to identify modifications in amino acid residues. We have found 106 different residues oxidatively damaged, being the main oxidized residues lysines, cysteines, arginines, prolines, histidines and tyrosines. Besides protein hydroxyl derivatives and oxygen additions, we detected other modifications such as deamidations, carbamylations and specific amino acid oxidative modifications. The oxidative damage preferentially affects particular subdomains of the protein at different time‐points. Results suggest the oxidative damage occurs first in exposed regions near cysteine disulfide bridges with residues like methionine, tryptophan, lysine, arginine, tyrosine and proline appearing as oxidatively modified. The damage extended afterwards with further oxidation of cysteine residues involved in disulfide bridges and other residues like histidine, phenylalanine and aspartic acid. The time‐course evaluation also shows the number of oxidized residues does not increase linearly, suggesting that oxidative unfolding of albumin occurs through a step‐ladder mechanism. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

15.
Unnatural amino acids extend the pharmacological formulator's toolkit. Strategies to prepare unnatural amino acid derivatives using Lewis acid‐activated allylsilane reactions are few. In this regard, we examined the utility of allylsilanes bearing an amino acid substituent in the reaction. Diastereoselective addition of methyl 2‐(N‐PG‐amino)‐3‐(trimethylsilyl)pent‐4‐enoate and methyl (E)‐2‐(N‐PG‐amino)‐3‐(trimethylsilyl)hex‐4‐enoate (PG=protecting group), 2 and 13 , respectively, to aromatic acetals in the presence of Lewis acids is described. Of those examined, TiCl4 was found to be the most effective Lewis acid for promoting the addition. At least 1 equiv. of TiCl4 was required to achieve high yields, whereas 2 equiv. of BF3?OEt2 were required for comparable outcomes. Excellent selectivity (>99% syn/anti) and high yield (up to 89%) were obtained with halo‐substituted aromatic acetals, while more electron‐rich electrophiles led to both lower yields and diastereoselectivities.  相似文献   

16.
We synthesized and carried out the conformational analysis of several hybrid dipeptides consisting of an α‐amino acid attached to a quaternary glyco‐β‐amino acid. In particular, we combined a S‐glycosylated β2,2‐amino acid and two different types of α‐amino acid, namely, aliphatic (alanine) and aromatic (phenylalanine and tryptophan) in the sequence of hybrid α/β‐dipeptides. The key step in the synthesis involved the ring‐opening reaction of a chiral cyclic sulfamidate, inserted in the peptidic sequence, with a sulfur‐containing nucleophile by using 1‐thio‐β‐D ‐glucopyranose derivatives. This reaction of glycosylation occurred with inversion of configuration at the quaternary center. The conformational behavior in aqueous solution of the peptide backbone and the glycosidic linkage for all synthesized hybrid glycopeptides was analyzed by using a protocol that combined NMR experiments and molecular dynamics with time‐averaged restraints (MD‐tar). Interestingly, the presence of the sulfur heteroatom at the quaternary center of the β‐amino acid induced θ torsional angles close to 180° (anti). Notably, this value changed to 60° (gauche) when the peptidic sequence displayed aromatic α‐amino acids due to the presence of CH–π interactions between the phenyl or indole ring and the methyl groups of the β‐amino acid unit.  相似文献   

17.
N‐Methyl β‐amino acids are generally required for application in the synthesis of potentially bioactive modified peptides and other oligomers. Previous work highlighted the reductive cleavage of 1,3‐oxazolidin‐5‐ones to synthesise N‐methyl α‐amino acids. Starting from α‐amino acids, two approaches were used to prepare the corresponding N‐methyl β‐amino acids. First, α‐amino acids were converted to N‐methyl α‐amino acids by the so‐called ‘1,3‐oxazolidin‐5‐one strategy’, and these were then homologated by the Arndt–Eistert procedure to afford N‐protected N‐methyl β‐amino acids derived from the 20 common α‐amino acids. These compounds were prepared in yields of 23–57% (relative to N‐methyl α‐amino acid). In a second approach, twelve N‐protected α‐amino acids could be directly homologated by the Arndt–Eistert procedure, and the resulting β‐amino acids were converted to the 1,3‐oxazinan‐6‐ones in 30–45% yield. Finally, reductive cleavage afforded the desired N‐methyl β‐amino acids in 41–63% yield. One sterically congested β‐amino acid, 3‐methyl‐3‐aminobutanoic acid, did give a high yield (95%) of the 1,3‐oxazinan‐6‐one ( 65 ), and subsequent reductive cleavage gave the corresponding AIBN‐derived N‐methyl β‐amino acid 61 in 71% yield (Scheme 2). Thus, our protocols allow the ready preparation of all N‐methyl β‐amino acids derived from the 20 proteinogenic α‐amino acids.  相似文献   

18.
It is a textbook knowledge that protein photoluminescence stems from the three aromatic amino acid residues of tryptophan(Trp), tyrosine (Tyr), and phenylalanine (Phe), with predominant contributions from Trp. Recently, inspired by the intrinsic emission of nonaromatic amino acids and poly(amino acids) in concentrated solutions and solids, we revisited protein light emission using bovine serum albumin (BSA) as a model. BSA is virtually nonemissive in dilute solutions (≤0.1 mg mL?1), but highly luminescent upon concentration or aggregation, showing unique concentration‐enhanced emission and aggregation‐induced emission (AIE) characteristics. Notably, apart from well‐documented UV luminescence, bright blue emission is clearly observed. Furthermore, persistent room‐temperature phosphorescence (p‐RTP) is achieved even in the amorphous solids under ambient conditions. This visible emission can be rationalized by the clustering‐triggered emission (CTE) mechanism. These findings not only provide an in‐depth understanding of the emissive properties of proteins, but also hold strong implications for further elucidating the basis of tissue autofluorescence.  相似文献   

19.
A catalytic strategy was developed for asymmetric substitution reactions at sp3‐hybridized carbon atoms by using a chiral alkylating agent generated in situ from trichloroacetimidate and a chiral phosphoric acid. The resulting chiral p‐methoxybenzyl phosphate selectively reacts with β‐amino alcohols rather than those without a β‐NH functionality. The use of an electronically and sterically tuned chiral phosphoric acid enables the kinetic resolution of amino alcohols through p‐methoxybenzylation with good enantioselectivity.  相似文献   

20.
A selected set of terminally protected β‐hexapeptides, each containing two nitroxide‐based (3R,4R)‐4‐amino‐1‐oxyl‐2,2,5,5‐tetramethylpyrrolidine‐3‐carboxylic acid (POAC) residues combined with four (1S,2S)‐2‐aminocyclopentane‐1‐carboxylic acid (ACPC) residues, was synthesised by using solution methods and was fully characterised. The two POAC residues are separated in the sequences by different numbers of intervening ACPC residues. The conformational features of the doubly spin‐labelled β‐hexapeptides were examined in chloroform by FTIR absorption and continuous‐wave electron paramagnetic resonance spectroscopic techniques. In particular, the biradical exchange coupling (J) between two POAC residues within each peptide indicates unambiguously that the secondary structure overwhelmingly adopted is the 12‐helix. Taken together, these results support the view that POAC is an excellent β‐amino acid for exploring this type of helical conformation in doubly labelled β‐peptides.  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