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 29 毫秒
1.
Representative members of a group of linear, N‐acylated polypeptide antibiotics (peptaibols) containing α‐aminoisobutyric acid (Aib) and, in part, isovaline (Iva), as well as proteinogenic amino acids and a C‐terminal‐bonded 2‐amino alcohol, were treated with anhydrous trifluoroacetic acid (TFA) at 37° for 0.5–26 h. The resulting fragments were separated by HPLC and characterized by electrospray ionization collision‐induced dissociation mass spectrometry (ESI‐CID‐MS). The following 16–20‐residue peptaibols were investigated: natural, microheterogeneous mixtures of antiamoebins and alamethicin F50, uniform paracelsin A, and synthetic trichotoxin A50/E. In the natural peptides, bonds formed between Aib (Iva) and Pro (Hyp) were rapidly and selectively cleaved within 0.5 h. Furthermore, TFA esters of the C‐terminal amino alcohols were formed. Depending on time, release of C‐terminal tri‐ and tetrapeptides as well as amino acids from the major fragments was observed. Synthetic homooligopeptides, namely Z‐ and Ac‐(Aib)10‐OtBu and Z‐(Aib)7‐OtBu, were analyzed for comparison. On treatment with TFA, a regular series of Z‐(Aib)10–5‐OH from Z‐(Aib)10‐OtBu were detected within 0.5 h, and, after 3 h, release of a regular series of Z‐(Aib)7–3‐OH from Z‐(Aib)7‐OtBu were observed. Moreover, concomitant release of the series of H‐(Aib)10–3‐OH from the decapeptide occurred. From these data, a repetitive cleavage mechanism via intermediate formation of C‐terminal oxazolones on trifluoroacetolysis is proposed. Furthermore, their formation and stability in native peptaibols are correlated with subtle structural differences in protein amino acids linked to Aib. From the conspicuous concordance of the formation and abundance of regular series of trifluoroacetolytic fragments and of positive ions of the b‐series in CID‐MS, the generation of intermediate oxazolonium ions in both gas and liquid phase is concluded.  相似文献   

2.
Enzymes implicated in cysteine and methionine metabolism such as cystathionine β‐lyase (CBL; EC 4.4.1.8), a pyridoxal‐5′‐phosphate (PLP)‐dependent carbon–sulfur lyase, have been shown to play a central role in the generation of sulfur compounds. This work describes the unprecedented cloning and characterization of the metC‐cystathionine β‐lyase from the axillary‐isolated strain Staphylococcus haemolyticus AX3, in order to determine its activity and its involvement in amino acid biosynthesis, and in the generation of sulfur compounds in human sweat. The gene contains a cysteine/methionine metabolism enzyme pattern, and also a sequence capable to effect β‐elimination. The recombinant enzyme was shown to cleave cystathionine into homocysteine and to convert methionine into methanethiol at low levels. No odor was generated after incubation of the recombinant enzyme with sterile human axillary secretions; sweat components were found to have an inhibitory effect. These results suggest that the generation of sulfur compounds by Staphylococci and the β‐lyase activity in human sweat are mediated by enzymes other than the metC gene or by the concerted activities of more than one enzyme.  相似文献   

3.
In this study, proteinogenic amino acids residues of dimeric dermorphin pentapeptides were replaced by the corresponding β3homo‐amino acids. The potency and selectivity of hybrid α/β dimeric dermorphin pentapeptides were evaluated by competetive receptor binding assay in the rat brain using [3H]DAMGO (a μ ligand) and [3H]DELT (a δ ligand). Tha analog containing β3homo‐Tyr in place of Tyr (Tyr‐d ‐Ala‐Phe‐Gly‐β3homo‐Tyr‐NH‐)2 showed good μ receptor affinity and selectivity (IC50 = 0.302, IC50 ratio μ/δ = 68) and enzymatic stability in human plasma. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
Insect cells are useful for the high‐yield production of recombinant proteins including chemokines and membrane proteins. In this study, we developed an insect cell‐based system for incorporating non‐natural amino acids into proteins at specific sites. Three types of promoter systems were constructed, and their efficiencies were compared for the expression of the prokaryotic amber suppressor tRNATyr in Drosophila melanogaster Schneider 2 cells. When paired with a variant of Escherichia coli tyrosyl‐tRNA synthetase specific for 3‐iodo‐L ‐tyrosine, the suppressor tRNA transcribed from the U6 promoter most efficiently incorporated the amino acid into proteins in the cells. The transient and stable introductions of these prokaryotic molecules into the insect cells were then compared in terms of the yield of proteins containing non‐natural amino acids, and the “transient” method generated a sevenfold higher yield. By this method, 4‐azido‐L ‐phenylalanine was incorporated into human interleukin‐8 at a specific site. The yield of the azido‐containing IL‐8 was 1 μg/1 mL cell culture, and the recombinant protein was successfully labeled with a fluorescent probe by the Staudinger–Bertozzi reaction.  相似文献   

5.
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   

6.
In recent years, there has been increased interest in bacterial methionine biosynthesis enzymes as antimicrobial targets because of their pivotal role in cell metabolism. C‐S lyase from Corynebacterium diphtheriae is a pyridoxal 5′‐phosphate‐dependent enzyme in the transsulfuration pathway that catalyzes the α,β‐elimination of sulfur‐containing amino acids, such as l ‐cystathionine, to generate ammonia, pyruvate, and homocysteine, the immediate precursor of L ‐methionine. In order to gain deeper insight into the functional and dynamic properties of the enzyme, mutants of two highly conserved active‐site residues, Y55F and Y114F, were characterized by UV‐visible absorbance, fluorescence, and CD spectroscopy in the absence and presence of substrates and substrate analogs, as well as by steady‐state kinetic studies. Substitution of Tyr55 with Phe apparently causes a 130‐fold decrease in at pH 8.5 providing evidence that Tyr55 plays a role in cofactor binding. Moreover, spectral data show that the mutant accumulates the external aldimine intermediate suggesting that the absence of interaction between the hydroxyl moiety and PLP‐binding residue Lys222 causes a decrease in the rate of substrate deprotonation. Mutation of Tyr114 with Phe slightly influences hydrolysis of l ‐cystathionine, and causes a change in substrate specificity towards l ‐serine and O‐acetyl‐l ‐serine compared to the wild type enzyme. These findings, together with computational data, provide useful insights in the substrate specificity of C‐S lyase, which seems to be regulated by active‐site architecture and by the specific conformation in which substrates are bound, and will aid in development of inhibitors. Proteins 2015; 83:78–90. © 2014 Wiley Periodicals, Inc.  相似文献   

7.
Sulfur‐containing aroma volatiles are important contributors to the distinctive aroma of melon and other fruits. Melon cultivars and accessions differ in the content of sulfur‐containing and other volatiles. l –methionine has been postulated to serve as a precursor of these volatiles. Incubation of melon fruit cubes with 13C‐ and 2H‐labeled l –methionine revealed two distinct catabolic routes into volatiles. One route apparently involves the action of an l ‐methionine aminotransferase and preserves the main carbon skeleton of l ‐methionine. The second route apparently involves the action of an l ‐methionine‐γ–lyase activity, releasing methanethiol, a backbone for formation of thiol‐derived aroma volatiles. Exogenous l ‐methionine also generated non‐sulfur volatiles by further metabolism of α–ketobutyrate, a product of l ‐methionine‐γ–lyase activity. α–Ketobutyrate was further metabolized into l –isoleucine and other important melon volatiles, including non‐sulfur branched and straight‐chain esters. Cell‐free extracts derived from ripe melon fruit exhibited l ‐methionine‐γ–lyase enzymatic activity. A melon gene (CmMGL) ectopically expressed in Escherichia coli, was shown to encode a protein possessing l ‐methionine‐γ–lyase enzymatic activity. Expression of CmMGL was relatively low in early stages of melon fruit development, but increased in the flesh of ripe fruits, depending on the cultivar tested. Moreover, the levels of expression of CmMGL in recombinant inbred lines co‐segregated with the levels of sulfur‐containing aroma volatiles enriched with +1 m/z unit and postulated to be produced via this route. Our results indicate that l ‐methionine is a precursor of both sulfur and non‐sulfur aroma volatiles in melon fruit.  相似文献   

8.
An Fe(II)/α‐ketoglutarate‐dependent dioxygenase, SadA, was obtained from Burkholderia ambifaria AMMD and heterologously expressed in Escherichia coli. Purified recombinant SadA had catalytic activity towards several N‐substituted l‐amino acids, which was especially strong with N‐succinyl l‐leucine. With the NMR and LC‐MS analysis, SadA converted N‐succinyl l‐leucine into N‐succinyl l‐threo‐β‐hydroxyleucine with >99% diastereoselectivity. SadA is the first enzyme catalysing β‐hydroxylation of aliphatic amino acid‐related substances and a potent biocatalyst for the preparation of optically active β‐hydroxy amino acids.  相似文献   

9.
Cleavage reactions at backbone loci are one of the consequences of oxidation of proteins and peptides. During α‐amidation, the Cα–N bond in the backbone is cleaved under formation of an N‐terminal peptide amide and a C‐terminal keto acyl peptide. On the basis of earlier works, a facilitation of α‐amidation by the thioether group of adjacent methionine side chains was proposed. This reaction was characterized by using benzoyl methionine and benzoyl alanyl methionine as peptide models. The decomposition of benzoylated amino acids (benzoyl‐methionine, benzoyl‐alanine, and benzoyl‐methionine sulfoxide) to benzamide in the presence of different carbohydrate compounds (reducing sugars, Amadori products, and reductones) was studied during incubation for up to 48 h at 80 °C in acetate‐buffered solution (pH 6.0). Small amounts of benzamide (0.3–1.5 mol%) were formed in the presence of all sugars and from all benzoylated species. However, benzamide formation was strongly enhanced, when benzoyl methionine was incubated in the presence of reductones and Amadori compounds (3.5–4.2 mol%). The reaction was found to be intramolecular, because α‐amidation of a similar 4‐methylbenzoylated amino acid was not enhanced in the presence of benzoyl‐methionine and carbohydrate compounds. In the peptide benzoyl‐alanyl‐methionine, α‐amidation at the methionine residue is preferred over α‐amidation at the benzoyl peptide bond. We propose here a mechanism for the enhancement of α‐amidation at methionine residues. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
The conformation of oligomers of β‐amino acids of the general type Ac‐[β‐Xaa]n‐NHMe (β‐Xaa = β‐Ala, β‐Aib, and β‐Abu; n = 1–4) was systematically examined at different levels of ab initio molecular orbital theory (HF/6‐31G*, HF/3‐21G). The solvent influence was considered employing two quantum‐mechanical self‐consistent reaction field models. The results show a wide variety of possibilities for the formation of characteristic elements of secondary structure in β‐peptides. Most of them can be derived from the monomer units of blocked β‐peptides with n = 1. The stability and geometries of the β‐peptide structures are considerably influenced by the side‐chain positions, by the configurations at the Cα‐ and Cβ‐atoms of the β‐amino acid constituents, and especially by environmental effects. Structure peculiarities of β‐peptides, in particular those of various helix alternatives, are discussed in relation to typical elements of secondary structure in α‐peptides. © 1999 John Wiley & Sons, Inc. Biopoly 50: 167–184, 1999  相似文献   

11.
β‐Amino acids containing hybrid peptides and β‐peptides show great potential as peptidomimetics. In this paper we describe the synthesis and affinity toward the µ‐ and δ‐opioid receptors of β‐peptides, analogues of Leu‐enkephalin, deltorphin I, dermorphin and α,β‐hybrides, analogues of deltorphin I. Substitution of α‐amino acid residues with β3homo‐amino acid residues, in general resulted in decrease of affinity to opioid receptors. However, the incorporation β3h‐D ‐Ala in position 2 or β3hPhe in position 3 of deltorphin I resulted in potent and selective ligand for δ‐opioid receptor. The NMR studies of β‐deltorphin I analogue suggest that conformational motions in the central part of the peptide backbone are partially restricted and some conformational preferences can be expected. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
Seven ascidiacyclamide [cyclo(–Ile–oxazoline–d ‐Val–thiazole–)2] (ASC) analogues incorporating the β‐amino acids βIle, βoxazoline, and/or d ‐βVal were synthesized. We then investigated the effects of the position and number of incorporated β‐amino acids on the structure, cytotoxicity, and copper binding by these seven analogues. The structural analyses revealed that both βIle and d ‐βVal favor a gauche‐type θ torsion angles, while βoxazoline favors a trans‐type θ torsion angle. Expansion of the macrocycle by incorporation of βIle or d ‐βVal readily induced molecular folding. On the other hand, the incorporation of two βoxazoline residues strongly extended the peptide conformation, and the incorporation of one was sufficient for the moderate restriction important for conformational equilibrium and cytotoxicity. Despite expansion of the macrocycles, the structure‐cytotoxicity relationships were largely maintained. In studies of complexation of the analogues with Cu (II) ion, the position and number of incorporated β‐amino acids had a large impact on the structure of the metal complex and may contribute to its stabilization.  相似文献   

13.
A series of tetrapeptide amides containing two aminoisobutyric acids (Aib) and two α‐methylphenylalanine ((αMe)Phe) units were prepared through the ‘azirine/oxazolone method’. New 2‐benzyl‐2‐methyl‐2H‐azirin‐3‐amines have been used for the selective introduction of (S)‐ and (R)‐(αMe)Phe, respectively. The solid‐state conformations of five tetrapeptide amides were determined by X‐ray crystallography. In all cases, two β‐turns stabilize 310‐helical conformations and it was confirmed that, in contrast to proteinogenic amino acids, the configuration of (αMe)Phe does not determine the screw sense of the helix.  相似文献   

14.
Optical resolution of β-(1-naphthyl)alanine and β-(2-naphthyl)alanine have been efficiently carried out through enzymatic hydrolysis of their methyl ester and/or N-acetyl ester derivatives by immobilized enzymes. Difficulties related to the lipophilic character of these amino acids were overcome by using emulsions of n-butyl acetate–water as reaction medium. The use of an automatic recirculating apparatus allowed reproducible and repetitive use of the immobilized biocatalysts.  相似文献   

15.
The construction of novel functional proteins has been a key area of protein engineering. However, there are few reports of functional proteins constructed from artificial scaffolds. Here, we have constructed a genetic library encoding α3β3 de novo proteins to generate novel scaffolds in smaller size using a binary combination of simplified hydrophobic and hydrophilic amino acid sets. To screen for folded de novo proteins, we used a GFP‐based screening system and successfully obtained the proteins from the colonies emitting the very bright fluorescence as a similar intensity of GFP. Proteins isolated from the very bright colonies (vTAJ) and bright colonies (wTAJ) were analyzed by circular dichroism (CD), 8‐anilino‐1‐naphthalenesulfonate (ANS) binding assay, and analytical size‐exclusion chromatography (SEC). CD studies revealed that vTAJ and wTAJ proteins had both α‐helix and β‐sheet structures with thermal stabilities. Moreover, the selected proteins demonstrated a variety of association states existing as monomer, dimer, and oligomer formation. The SEC and ANS binding assays revealed that vTAJ proteins tend to be a characteristic of the folded protein, but not in a molten‐globule state. A vTAJ protein, vTAJ13, which has a packed globular structure and exists as a monomer, was further analyzed by nuclear magnetic resonance. NOE connectivities between backbone signals of vTAJ13 suggested that the protein contains three α‐helices and three β‐strands as intended by its design. Thus, it would appear that artificially generated α3β3 de novo proteins isolated from very bright colonies using the GFP fusion system exhibit excellent properties similar to folded proteins and would be available as artificial scaffolds to generate functional proteins with catalytic and ligand binding properties.  相似文献   

16.
V L Richmond 《Life sciences》1986,39(3):263-268
Methionine, an essential amino acid, and cysteine are the major sulfur-containing amino acids in the body and both are thought to be synthesized predominantly in plants and micro-organisms. Methylsulfonylmethane (MSM) is a natural constituent of the environment in which it is found in plants, in milk and urine of both bovines and humans, is a normal oxidation product of dimethyl sulfoxide (DMSO) also in the natural environment and may be part of the natural global sulfur cycle. To determine whether sulfur from methylsulfonylmethane (MSM) is incorporated into sulfur amino acids, I fed 35S-MSM to guinea pigs. 35S was incorporated into peptidyl methionine and cysteine of guinea pig serum proteins. The specific activity of 35S-methionine was 30% greater than for 35S-cysteine, suggesting a precursor-product relationship. Total specific activity of serum proteins was increased by only 30% with a 100% increase of administered 35S-MSM, suggesting a limiting step in synthesis. Approximately 1% of the radioactivity was recovered in serum proteins, none in the feces and most was excreted in the urine. Microorganisms of intestinal lumen may be responsible for the incorporation of the 35S of MSM into sulfur amino acids. MSM may provide a source of sulfur for essential animal methionine by mechanisms not yet elucidated in either animals or micro-organisms.  相似文献   

17.
β‐Amino acids containing α,β‐hybrid peptides show great potential as peptidomimetics. In this paper, we describe the synthesis and affinity to μ‐opioid and δ‐opioid receptors of α,β‐hybrids, analogs of the tetrapeptide Tyr‐ d ‐Ala‐Phe‐Phe‐NH2 (TAPP). Each amino acid was replaced with an l ‐ or d ‐β3h‐amino acid. All α,β‐hybrids of TAPP analogs were synthesized in solution and tested for affinity to μ‐opioid and δ‐opioid receptors. The analog Tyr‐β3h‐ d ‐Ala‐Phe‐PheNH2 was found to be as active as the native tetrapeptide. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
Stable peptides have been explored as epitope mimics for protein–protein and protein–nucleic acid interactions; however, presentation of a regular structure is critical. Aromatic interactions are ubiquitous and are competent at stabilizing a β‐hairpin fold. The greatest stabilization has been reported from pairs of tryptophan side chains. Naphthylalanine residues are often used as tryptophan replacements, but it is not clear if 1‐naphthylalanine or 2‐naphthylalanine is adequate at replicating the geometry and stability observed with tryptophan aromatic interactions. Herein, a 12‐residue peptide has been constructed with laterally disposed aromatic amino acids. A direct comparison is made between tryptophan and other bicyclic, unnatural amino acids. Significant stabilization is gained from all bicyclic amino acids; however, geometric analysis shows that only 1‐naphthylalanine adopts a similar edge to face geometry as tryptophan, whereas the 2‐naphthylalanine appears most similar to a substituted phenylalanine. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
Poly‐γ‐glutamate (γ‐PGA) has applications in food, medical, cosmetic, animal feed, and wastewater industries. Bacillus subtilis DB430, which possesses the γ‐PGA synthesis ywsC‐ywtAB genes in its chromosome, cannot produce γ‐PGA. An efficient synthetic expression control sequence (SECS) was introduced into the upstream region of the ywtABC genes, and this resulted in γ‐PGA‐producing B. subtilis mutant strains. Mutant B. subtilis PGA6‐2 stably produces high levels of γ‐PGA in medium A without supplementation of extra glutamic acid or ammonium chloride. The mutant B. subtilis PGA 6‐2 is not only a γ‐PGA producer, but it is also a candidate for the genetic and metabolic engineering of γ‐PGA production. © 2010 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

20.
Both Type I' and Type II' β‐turns have the same sense of the β‐turn twist that is compatible with the β‐sheet twist. They occur predominantly in two residue β‐hairpins, but the occurrence of Type I' β‐turns is two times higher than Type II' β‐turns. This suggests that Type I' β‐turns may be more stable than Type II' β‐turns, and Type I' β‐turn sequence and structure can be more favorable for protein folding than Type II' β‐turns. Here, we redesigned the native Type II' β‐turn in GFP to Type I' β‐turn, and investigated its effect on protein folding and stability. The Type I' β‐turns were designed based on the statistical analysis of residues in natural Type I' β‐turns. The substitution of the native “GD” sequence of i+1 and i+2 residues with Type I' preferred “(N/D)G” sequence motif increased the folding rate by 50% and slightly improved the thermodynamic stability. Despite the enhancement of in vitro refolding kinetics and stability of the redesigned mutants, they showed poor soluble expression level compared to wild type. To overcome this problem, i and i + 3 residues of the designed Type I' β‐turn were further engineered. The mutation of Thr to Lys at i + 3 could restore the in vivo soluble expression of the Type I' mutant. This study indicates that Type II' β‐turns in natural β‐hairpins can be further optimized by converting the sequence to Type I'. Proteins 2014; 82:2812–2822. © 2014 Wiley Periodicals, Inc.  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