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 359 毫秒
1.
Carbodiimide-mediated peptide synthesis in aqueous solution has been studied with respect to self-ordering of amino acids. The copolymerisation of amino acids in the presence of glutamic acid or pyroglutamic acid leads to short pyroglutamyl peptides. Without pyroglutamic acid the formation of higher polymers is favoured.The interactions of the amino acids and the peptides, however, are very complex. Therefore, the experimental results are rather difficult to explain. Some of the experimental results, however, can be explained with the aid of computer simulation programs. Regarding only the tripeptide fraction the copolymerisation of pyroGlu, Ala and Leu, as well as the simulated copolymerisation lead to pyroGlu-Ala-Leu as the main reaction product. The amino acid composition of the insoluble peptides formed during the copolymerisation of Ser, Gly, Ala, Val, Phe, Leu and Ile corresponds in part to the computer-simulated copolymerisation data.  相似文献   

2.
Carbodiimide-mediated peptide synthesis in aqueous solution at room temperature has been studied with respect to self-ordering of amino acids. Inasmuch as glutamic acid is readily converted into pyroglutamic acid, the peptides formed by copolymerization of glutamic acid with other amino acids are preferentially pyroglutamyl-peptides. Competition experiments were carried out to determine the influence of the amino acid side chains on the reactivity of amino acids and peptides during the dehydration condensation. The results show that the self-ordering process is controlled by both the activated carboxyl component (amino acid or growing peptide) and the incoming amino acid. A condensation of pyroglutamic acid, alanine and another amino acid component Xxx (Xxx = Gly, Val, Leu or Gly-Tyr) preferentially yielded the dipeptide pyroGlu-Ala, but the formation of the tripeptide pyroGlu-Ala-Ala became strongly reduced because of competing reactions. A simple explanation for the observed selectivities is not at hand. Polypeptides were so far only obtained when they were allowed to precipitate in the reaction system. Evidence for the non-random copolymerization of larger peptides is presented as well.  相似文献   

3.
P Melius  W L Hubbard 《Bio Systems》1987,20(3):213-217
It has been established indirectly that the N-termini of the thermal polyamino acids are pyroglutamic acid. This was determined by trifluoroacetic acid hydrolysis of the lactam ring followed by Dansyl labelling. The polyamino acids contained Ala, Gly, Glu, Leu, Phe, and Pro. In the experiments described here, the presence of pyroglutamic acid at the N-terminus of a polyamino acid was determined directly by the use of pyrrolidone carboxylyl peptidase. The enzyme catalyzes the removal of pyroglutamyl residues at the N-terminus of polypeptide chains. The polyamino acids used in these studies contained glutamic acid, aspartic acid, alanine, glycine, isoleucine, proline and valine. Alkaline hydrolysis was also used to determine indirectly that the N-termini of these polyamino acids are pyroglutamic acid. Another interesting finding was that many of the amino acids in the polymerization mixture were found to occur penultimate to the N-terminal amino acid. This is interpreted to mean that the diffusible fraction contains many polyamino acids.  相似文献   

4.
The effect of single amino acid substitutions at positions 18 and 20 on the T-cell determinant (TD) character of peptide p12-26 from lambda repressor protein and on its recognition by a monoclonal antibody was studied by means of 40 synthetic peptides of a length of 15 amino acids. ELISA competition experiments showed that the identity of amino acid at position 20 is very important for antibody recognition, whereas that of amino acid at position 18 is much less important. In contrast, both Leu 18 and Ala 20 are important residues in defining the TD character of peptide p12-26. The most tolerated replacements, ordered in increasing disrupting power are: Ala 20 by Cys, Ser or Gly and Leu 18 by Ile or Val. Any other amino acid replacement completely abolishes the TD capacity of peptide p12-26. The peptides used in this study were synthesized using a multiple solid-phase peptide synthesizer newly designed. Their purity was very high as shown by amino acid sequence experiments.  相似文献   

5.
Helix propensities of the amino acids have been measured in alanine-based peptides in the absence of helix-stabilizing side-chain interactions. Fifty-eight peptides have been studied. A modified form of the Lifson-Roig theory for the helix-coil transition, which includes helix capping (Doig AJ, Chakrabartty A, Klingler TM, Baldwin RL, 1994, Biochemistry 33:3396-3403), was used to analyze the results. Substitutions were made at various positions of homologous helical peptides. Helix-capping interactions were found to contribute to helix stability, even when the substitution site was not at the end of the peptide. Analysis of our data with the original Lifson-Roig theory, which neglects capping effects, does not produce as good a fit to the experimental data as does analysis with the modified Lifson-Roig theory. At 0 degrees C, Ala is a strong helix former, Leu and Arg are helix-indifferent, and all other amino acids are helix breakers of varying severity. Because Ala has a small side chain that cannot interact significantly with other side chains, helix formation by Ala is stabilized predominantly by the backbone ("peptide H-bonds"). The implication for protein folding is that formation of peptide H-bonds can largely offset the unfavorable entropy change caused by fixing the peptide backbone. The helix propensities of most amino acids oppose folding; consequently, the majority of isolated helices derived from proteins are unstable, unless specific side-chain interactions stabilize them.  相似文献   

6.
Theil R  Scheit KH 《The EMBO journal》1983,2(7):1159-1163
Analytical ultracentrifugation of highly purified seminalplasmin revealed a molecular mass of 6300. Amino acid analysis of the protein preparation indicated the absence of sulfur-containing amino acids cysteine and methionine. The amino acid sequence of seminalplasmin was determined by manual Edman degradation of peptides obtained by proteolytic enzymes trypsin, chymotrypsin and thermolysin: NH2-Ser Asp Glu Lys Ala Ser Pro Asp Lys His His Arg Phe Ser Leu Ser Arg Tyr Ala Lys Leu Ala Asn Arg Leu Ser Lys Trp Ile Gly Asn Arg Gly Asn Arg Leu Ala Asn Pro Lys Leu Leu Glu Thr Phe Lys Ser Val-COOH. The number of amino acids according to the sequence were 48, the molecular mass 6385. As predicted from the sequence, seminalplasmin very likely contains two α-helical domains in which residues 8-17 and 40-48 are involved. No evidence for the existence of β-sheet structures was obtained. Treatment of seminalplasmin with the above proteases as well as with amino peptidase M and carboxypeptidase Y completely eliminated biological activity.  相似文献   

7.
Yuko Okamoto 《Proteins》1994,19(1):14-23
Monte Carlo simulated annealing is applied to the study of the α-helix-forming tendencies of seven nonpolar amino acids, Ala, Leu, Met, Phe, Ile, Val, and Gly. Homooligomers of 10 amino acids are used and the helix tendency is calculated by folding α-helicies from completely random initial conformations. The results of the simulation imply that Met, Ala, and Leu are helix formers and that Val, Ile, and Gly are helix breakers, while Phe comes in between the two groups. The differences between helix formers and breakers turned out to be large in agreement with the recent experiments with short peptides. It is argued from the energy distributions of the obtained conformations that the helix tendency is small for the helix breakers because of steric hindrance of side chains. Homoglycine is shown to favor a random coil conformation. The β-strand tendencies of the same homooligomers are also considered, and they are shown to agree with the frequencies of amino acids in β-sheet from the protein data base. © 1994 Wiley-Liss, Inc.  相似文献   

8.
The genomic as well as structural relationship of phycobiliproteins (PBPs) in different cyanobacterial species are determined by nucleotides as well as amino acid composition. The genomic GC constituents influence the amino acid variability and codon usage of particular subunit of PBPs. We have analyzed 11 cyanobacterial species to explore the variation of amino acids and causal relationship between GC constituents and codon usage. The study at the first, second and third levels of GC content showed relatively more amino acid variability on the levels of G3 + C3 position in comparison to the first and second positions. The amino acid encoded GC rich level including G rich and C rich or both correlate the codon variability and amino acid availability. The fluctuation in amino acids such as Arg, Ala, His, Asp, Gly, Leu and Glu in α and β subunits was observed at G1C1 position; however, fluctuation in other amino acids such as Ser, Thr, Cys and Trp was observed at G2C2 position. The coding selection pressure of amino acids such as Ala, Thr, Tyr, Asp, Gly, Ile, Leu, Asn, and Ser in α and β subunits of PBPs was more elaborated at G3C3 position. In this study, we observed that each subunit of PBPs is codon specific for particular amino acid. These results suggest that genomic constraint linked with GC constituents selects the codon for particular amino acids and furthermore, the codon level study may be a novel approach to explore many problems associated with genomics and proteomics of cyanobacteria.  相似文献   

9.
Straight-chain non-polar amino acids are good helix-formers in water   总被引:6,自引:0,他引:6  
For comparison with earlier data on naturally occurring non-polar amino acids (Ala, Leu, Phe, Val, Ile), the comparative helix-forming tendencies have been measured for non-polar amino acid residues that have unbranched side-chains, with an ethyl, propyl or butyl group, and also for methionine. The substitutions are made in a 17-residue alanine-based peptide. The results show that straight-chain non-polar amino acids have high helix-forming tendencies compared to beta-branched non-polar amino acids. Restriction of side-chain conformations in the helix, with a corresponding reduction in conformational entropy, is the likely explanation. There is a small increase in helix-forming tendency as the side-chain increases in length from ethyl to butyl, which suggests that a helix-stabilizing hydrophobic interaction is being detected.  相似文献   

10.
The mechanism of the thermal polymerization (at 180°C) of glutamic acid, tyrosine, and glycine has been studied. Glutamic acid is quickly and almost completely converted into pyroglutamic acid. The only dipeptide that is formed by dimerization of the remaining two amino acids is cyclic glycyl-tyrosine (a diketopiperazin). In a secondary reaction pyroglutamic acid interacts with cyclic glycyl-tyrosine and yields pyroglutamyl-glycyltyrosine and pyroglutamyl-tyrosyl-glycine. Other di- or tripeptides are not observed. The preferential appearance of the two pyroglutamyl-peptides has been reported earlier by Nakashima et al. (1977). The present data explain those results. Model experiments show that cyclic glycyl-tyrosine can also be cleaved by other acids or bases. In the presence of acetic acid at 118°C N-acetyl-glycyl-tyrosine is the major product. Partial hydrolysis predominantly yields tyrosyl-glycine. These effects are explained by stereospecific interactions. The results on self-ordering of amino acids during peptide formation are discussed in respect of the origin of prebiotic enzymes and genetic information.  相似文献   

11.
Based on the published bovine DNase sequence (Liao, T.-H., Salnikow, J., Moore, S., and Stein, W. H. (1973) J. Biol. Chem. 248, 1489-1495), the ovine DNase sequence is derived from the amino acid compositions of isolated short peptides covering all regions of the intact polypeptide. The sequence is substantiated by results of automated Edman degradation of the intact polypeptide and of the two middle CNBr fragments, and by elucidation of the complete sequence of the COOH-terminal CNBr peptide. The 12 changes from bovine to ovine DNase are at residues 22 (Ala to Ser), 29 (Val to Leu), 35 (Val to Ala), 54 (Tyr to Asp), 62 (Thr to Ser), 83 (Leu to Val), 121 (His to Pro), 127 (Glu to Ala), 132 (Ala to Pro), 159 (His to Asp), 163 (Val to Ile), and 231 (Ala to Val). A minor genetic variant form of ovine DNase has Val at residue 163. The data from automated Edman degradation of the largest CNBr peptide of bovine DNase show that the published bovine DNase sequence is in error and that an Ile-Val-Arg tripeptide must be inserted between Arg-27 and Arg-28. The corrected sequence is substantiated by two peptides covering this region each with three amino acids more than the published sequence. Comparison of the bovine, ovine, and porcine DNase sequences reveals the following: with the revised bovine sequence, all three DNase sequences can be aligned without a gap; all three DNases have a carbohydrate side chain at Asn-18, but only porcine DNase has carbohydrate at Asn-106; there are 12 changes between bovine and ovine DNases, 56 between bovine and porcine, and 50 between ovine and porcine; there are six highly variable regions and four invariable ones; bovine and ovine DNases have the same length while porcine DNase is longer by 2 amino acid residues at the COOH terminus; the residues around the nucleotide-binding site, the four pairs of salt bridges, and the essential His-134 groups are not changed.  相似文献   

12.
Avrahami D  Oren Z  Shai Y 《Biochemistry》2001,40(42):12591-12603
The initial stages leading to the binding and functioning of membrane-active polypeptides including hormones, signal sequences, and lytic peptides are mainly governed by electrostatic attraction and hydrophobic partitioning between water and lipid bilayers. Antimicrobial peptides serve as an important model for studying the details of these initial steps. However, a systematic analysis of the contribution of multiple hydrophobic amino acids to these steps have been hindered by the propensity of many peptides to aggregate and become inactivated in solution. To this end, we synthesized a series of model amphipathic all L-amino acid peptides and their diastereomers with the sequence KX(3)KWX(2)KX(2)K, where X = Gly, Ala, Val, Ile, or Leu. The effect of the aliphatic amino acids on the biological activity, binding, structure, membrane localization, and mode of action of these peptides was investigated. Most of the L-amino acid peptides oligomerized and adopted distinct structures in solution and in a membrane mimetic environment. Among this group only the Leu containing peptide was hemolytic and highly active on most bacteria tested. The Val- and Leu-containing peptides were hemolytic but inactive toward most bacteria tested. In contrast, the diastereomeric peptides were monomeric and unstructured in solution, but they adopted distinct structures upon membrane binding. While hemolytic activity was drastically reduced, the spectrum of antibacterial activity was preserved or increased. Importantly, we found a direct correlation with the diastereomers between hydrophobicity and propensity to form a helical/distorted-helix and activity (induced membrane leakage and antibacterial activity), despite the fact that they contained 30% D-amino acids. Furthermore, efficient increase in membrane permeability can proceed through different mechanisms. Specifically, the Leu-containing diastereomeric peptide micellized vesicles and possibly bacterial membranes while the Ile-containing diastereomeric peptide fused model membranes and irregularly disrupted bacterial membranes.  相似文献   

13.
Presecretory signal peptides of 39 proteins from diverse prokaryotic and eukaryotic sources have been compared. Although varying in length and amino acid composition, the labile peptides share a hydrophobic core of approximately 12 amino acids. A positively charged residue (Lys or Arg) usually precedes the hydrophobic core. Core termination is defined by the occurrence of a charged residue, a sequence of residues which may induce a beta-turn in a polypeptide, or an interruption in potential alpha-helix or beta-extended strand structure. The hydrophobic cores contain, by weight average, 37% Leu: 15% Ala: 10% Val: 10% Phe: 7% Ile plus 21% other hydrophobic amino acids arranged in a non-random sequence. Following the hydrophobic cores (aligned by their last residue) a highly non-random and localized distribution of Ala is apparent within the initial eight positions following the core: (formula; see text) Coincident with this observation, Ala-X-Ala is the most frequent sequence preceding signal peptidase cleavage. We propose the existence of a signal peptidase recognition sequence A-X-B with the preferred cleavage site located after the sixth amino acid following the core sequence. Twenty-two of the above 27 underlined Ala residues would participate as A or B in peptidase cleavage. Position A includes the larger aliphatic amino acids, Leu, Val and Ile, as well as the residues already found at B (principally Ala, Gly and Ser). Since a preferred cleavage site can be discerned from carboxyl and not amino terminal alignment of the hydrophobic cores it is proposed that the carboxyl ends are oriented inward toward the lumen of the endoplasmic reticulum where cleavage is thought to occur. This orientation coupled with the predicted beta-turn typically found between the core and the cleavage site implies reverse hairpin insertion of the signal sequence. The structural features which we describe should help identify signal peptides and cleavage sites in presumptive amino acid sequences derived from DNA sequences.  相似文献   

14.
The interaction of amino acid residues with polyribonucleotides was characterized by measurements of melting temperatures (tm) for poly(A).poly(U) and poly(I).poly(C) as functions of the concentrations of various amino acid amides. The amides of hydrophilic amino acids lead to a continuous increase of tm with increasing concentration, whereas amides of hydrophobic amino acids induce a decrease of tm at low concentrations (approximately 1 mM) followed by an increase at higher concentrations. Analysis of the data by a simple site model provides the affinity of each ligand for the double helix relative to that for the single strands. This parameter decreases in the order Ala greater than Gly greater than Ser greater than Asn greater than Pro greater than Met, Val greater than Ile, Leu for poly(A).poly(U) and Ala, Gly, Ser greater than Asn greater than Pro greater than Val greater than Ile, Met, Leu for poly(I).poly(C). The special effects of hydrophobic amino acids may be related to the similarity of the codons for these amino acids. A simple model for assignment of codons to amino acids is proposed.  相似文献   

15.
The trans-inhibition potency of intracellular amino acids on the transport of various amino acids follows the same sequence,viz. Pro(Lys), Phe, Glu, Ala, Gly, Leu, and α-aminoisobutyric acid. The same sequence was found for the reciprocal of trans-inhibition constants. It appears that the intracellular amino acid itself or a derivative thereof acts on a component that is common to all amino acid transport systems of baker’s yeast.  相似文献   

16.
Complete amino acid sequence of staphylococcal enterotoxin A   总被引:10,自引:0,他引:10  
The amino acid sequence of staphylococcal enterotoxin A is presented. Staphylococcal enterotoxin A is a single-chain polypeptide which consists of 233 amino acid residues with a molecular weight of 27,078 and has the amino acid composition Cys2, Asp17, Asn19, Thr16, Ser13, Glu15, Gln12, Pro4, Gly15, Ala7, Val13, Met2, Ile10, Leu23, Tyr18, Phe8, His6, Lys24, Arg7, Trp2, with serine as both amino- and carboxyl-terminal amino acids. Automated sequence analysis of intact enterotoxin A, as well as characterization of the peptides obtained from cyanogen bromide treatment and trypsin and chymotrypsin digestion, led to the elucidation of the complete primary structure of this protein. Less structural homology is observed among staphylococcal enterotoxins A, B (Huang, I-Y., and Bergdoll, M. S. (1970) J. Biol. Chem. 245, 3518-3525), and C1 (Schmidt, J. J., and Spero, L. (1983) J. Biol. Chem. 258, 6300-6306) than that seen between enterotoxins B and C1.  相似文献   

17.
Accurate random coil alpha-proton chemical shift values are essential for precise protein structure analysis using chemical shift index (CSI) calculations. The current study determines the chemical shift effects of pH, urea, peptide length and neighboring amino acids on the alpha-proton of Ala using model peptides of the general sequence GnXaaAYaaGn, where Xaa and Yaa are Leu, Val, Phe, Tyr, His, Trp or Pro, and n = 1-3. Changes in pH (2-6), urea (0-1M), and peptide length (n = 1-3) had no effect on Ala alpha-proton chemical shifts. Denaturing concentrations of urea (8M) caused significant downfield shifts (0.10 +/- 0.01 ppm) relative to an external DSS reference. Neighboring aliphatic residues (Leu, Val) had no effect, whereas aromatic amino acids (Phe, Tyr, His and Trp) and Pro caused significant shifts in the alanine alpha-proton, with the extent of the shifts dependent on the nature and position of the amino acid. Smaller aromatic residues (Phe, Tyr, His) caused larger shift effects when present in the C-terminal position (approximately 0.10 vs. 0.05 ppm N-terminal), and the larger aromatic tryptophan caused greater effects in the N-terminal position (0.15 ppm vs. 0.10 C-terminal). Proline affected both significant upfield (0.06 ppm, N-terminal) and downfield (0.25 ppm, C-terminal) chemical shifts. These new Ala correction factors detail the magnitude and range of variation in environmental chemical shift effects, in addition to providing insight into the molecular level interactions that govern protein folding.  相似文献   

18.
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   

19.
The catalytic effect of activated alumina on amino acid condensation was investigated. The readiness of amino acids to form peptide sequences was estimated on the basis of the yield of dipeptides and was found to decrease in the order glycine (Gly), alanine (Ala), leucine (Leu), valine (Val), proline (Pro). For example, approximately 15% Gly was converted to the dipeptide (Gly(2)), 5% to cyclic anhydride (cyc(Gly(2))) and small amounts of tri- (Gly(3)) and tetrapeptide (Gly(4)) were formed after 28 days. On the other hand, only trace amounts of Pro(2) were formed from proline under the same conditions. Preferential formation of certain sequences was observed in the mixed reaction systems containing two amino acids. For example, almost ten times more Gly-Val than Val-Gly was formed in the Gly+Val reaction system. The preferred sequences can be explained on the basis of an inductive effect that side groups have on the nucleophilicity and electrophilicity, respectively, of the amino and carboxyl groups. A comparison with published data of amino acid reactions in other reaction systems revealed that the main trends of preferential sequence formation were the same as those described for the salt-induced peptide formation (SIPF) reaction. The results of this work and other previously published papers show that alumina and related mineral surfaces might have played a crucial role in the prebiotic formation of the first peptides on the primitive earth.  相似文献   

20.
The mammalian Na(+)/H(+) exchanger isoform 1 (NHE1) resides on the plasma membrane and exchanges one intracellular H(+) for one extracellular Na(+). It maintains intracellular pH and regulates cell volume, and cell functions including growth and cell differentiation. Previous structural and functional studies on TMVI revealed several amino acids that are potentially pore lining. We examined these and other critical residues by site-directed mutagenesis substituting Asn227→Ala, Asp, Arg; Ile233→Ala; Leu243→Ala; Glu247→Asp, Gln; Glu248→Asp, Gln. Mutant NHE1 proteins were characterized in AP-1 cells, which do not express endogenous NHE1. All the TMVI critical amino acids were highly sensitive to substitution and changes often lead to a dysfunctional protein. Mutations of Asn227→Ala, Asp, Arg; Ile233→Ala; Leu243→Ala; Glu247→Asp; Glu248→Gln yielded significant reduction in NHE1 activity. Mutants of Asn227 demonstrated defects in protein expression, targeting and activity. Substituting Asn227→Arg and Ile233→Ala decreased the surface localization and expression of NHE1 respectively. The pore lining amino acids Ile233 and Leu243 were both essential for activity. Glu247 was not essential, but the size of the residue at this location was important while the charge on residue Glu248 was more critical to NHE1 function. Limited trypsin digestion on Leu243→Ala and Glu248→Gln revealed that they had increased susceptibility to proteolytic attack, indicating an alteration in protein conformation. Modeling of TMVI with TMXI suggests that these TM segments form part of the critical fold of NHE1 with Ile233 and Leu465 of TMXI forming a critical part of the extracellular facing ion conductance pathway.  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