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.
Peptide chemical ligation chemistries, which allow the chemoselective coupling of unprotected peptide fragments, are useful tools for synthesizing native polypeptides or unnatural peptide‐based macromolecules. We show here that the phenylthiocarbonyl group can be easily introduced into peptides on α or ε amino groups using phenylthiochloroformate and standard solid‐phase method. It reacts chemoselectively with cysteinyl peptides to give an alkylthiocarbamate bond. S,N‐shift of the alkylaminocarbonyl group from the Cys side chain to the α‐amino group did not occur. The method was used for linking two peptide chains through their N‐termini, for the synthesis of a cyclic peptide or for the synthesis of di‐ or tetravalent multiple antigenic peptides (MAPs). Thiocarbamate ligation is thus complementary to thioether, thioester or disulfide ligation methods. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
Various peptides and derivatives of peptides and amino acids were synthesized and tasted, systematically, to elucidate the relationship between bitterness and chemical structures of peptides.

We have found that: 1. Peptides become more bitter than the original amino acids when their amino and carboxyl groups are blocked and when peptide bond is formed. 2. A peptide molecule with a high content of amino acids with hydrophobic side chains will develop bitter taste. 3. The amino acids in a peptide chain independently contribute to bitterness regardless of amino acid sequences and configuration.  相似文献   

3.
Incorporation of ferrocenyl group to peptides is an efficient method to alter their hydrophobicity. Ferrocenyl group can also act as an electrochemical probe when incorporated onto functional peptides. Most often, ferrocene is incorporated onto peptides post‐synthesis via amide, ester or triazole linkages. Stable amino acids containing ferrocene as a C‐linked side chain are potentially useful building units for the synthesis of ferrocene‐containing peptides. We report here an efficient route to synthesize ferrocene‐containing amino acids that are stable and can be used in peptide synthesis. Coupling of 2‐ferrocenyl‐1,3‐dithiane and iodides derived from aspartic acid or glutamic acid using n‐butyllithium leads to the incorporation of a ferrocenyl unit to the δ‐position or ε‐position of an α‐amino acid. The reduction or hydrolysis of the dithiane group yields an alkyl or an oxo derivative. The usability of the synthesized amino acids is demonstrated by incorporating one of the amino acids in both C‐terminus and N‐terminus of tripeptides in solution phase. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
Some amino acids are particularly susceptible to degradation in long-lived proteins. Foremost among these are asparagine, aspartic acid and serine. In the case of serine residues, cleavage of the peptide bond on the N-terminal side, as well as racemisation, has been observed. To investigate the role of the hydroxyl group, and whether cleavage and racemisation are linked by a common mechanism, serine peptides with a free hydroxyl group were compared to analogous peptides where the serine hydroxyl group was methylated. Peptide bond cleavage adjacent to serine was increased when the hydroxyl group was present, and this was particularly noticeable when it was present as the hydroxide ion. Adjacent amino acid residues also had a pronounced affect on cleavage at basic pH, with the SerPro motif being especially susceptible to scission. Methylation of the serine hydroxyl group abolished truncation, as did insertion of a bulky amino acid on the N-terminal side of serine. By contrast, racemisation of serine occurred to a similar extent in both O-methylated and unmodified peptides. On the basis of these data, it appears that racemisation of Ser, and cleavage adjacent to serine, occur via separate mechanisms. Addition of water across the double bond of dehydroalanine was not detected, suggesting that this mechanism was unlikely to be responsible for conversion of l-serine to d-serine. Abstraction of the alpha proton may account for the majority of racemisation of serine in proteins.  相似文献   

5.
Hydrocarbon stapling of peptides is a powerful technique to transform linear peptides into cell-permeable helical structures that can bind to specific biological targets. In this study, we have used high resolution solution NMR techniques complemented by dynamic light scattering to characterize extensively a family of hydrocarbon stapled peptides with known inhibitory activity against HIV-1 capsid assembly to evaluate the various factors that modulate activity. The helical peptides share a common binding motif but differ in charge, the length, and position of the staple. An important outcome of the study was to show the peptides, share a propensity to self-associate into organized polymeric structures mediated predominantly by hydrophobic interactions between the olefinic chain and the aromatic side-chains from the peptide. We have also investigated in detail the structural significance of the length and position of the staple, and of olefinic bond isomerization in stabilizing the helical conformation of the peptides as potential factors driving polymerization. This study presents the numerous challenges of designing biologically active stapled peptides and the conclusions have broad implications for optimizing a promising new class of compounds in drug discovery.  相似文献   

6.
Native chemical ligation of unprotected peptides in organic solvents has been previously reported as a fast, efficient, and suitable method for coupling of hydrophobic peptides. However, it has not been determined whether the reaction can be carried out without possible side reactions or racemization. Here, we present a study on the chemoselectivity of this method by model reactions designed to test the reactivity of Arg and Lys side chains as well as that of α‐amino groups. A possible racemization of the C‐terminal amino acid of the N‐terminal peptide was also investigated. The results show that ligation in organic solvents can be conducted chemoselectively without side reactions with other nucleophilic groups. Furthermore, no racemization of the C‐terminal amino acid was observed if both educts were added simultaneously. Thus, native chemical ligation can be performed either in aqueous buffer systems or in organic solvents paving the way for the synthesis of larger hydrophobic peptides and/or membrane proteins. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

7.
The molecular conformations of the linear oligopeptides H-(L -Ala)n-L -Pro-OH, with n = 1,2 and 3, have been investigated. 13C nmr observation of the equilibrium between the cis and trans forms of the Ala-Pro peptide bond indicated the occurrence of nonrandom conformations in solutions of these flexible peptides. The formation of the nonrandom species containing the cis form of the Ala-Pro bond was found to depend on the deprotonation of the carboxylic acid group of proline, the solvent, and the ionic strength in aqueous solution. The influence of intramolecular hydrogen bonding on the relative conformational energies of the species containing the cis and trans Ala-Pro peptide bond was studied by comparison of the peptides H-(Ala)n-Pro-OH with analogous molecules where hydrogen bond formation was excluded by the covalent structure. In earlier work a hydrogen bond between the protonated terminal carboxylic acid group and the carbonyl oxygen of the penultimate amino acid residue had been suggested to stabilize conformations including trans proline. For the systems described here this hypothesis can be ruled out, since the cis:trans ratio is identical for molecules with methyl ester protected and free protonated terminal carboxylic acid groups of proline. Direct evidence for hydrogen bond formation between the deprotonated terminal carboxylic acid group and the amide proton of the penultimate amino acid residue in the molecular species containing cis proline was obtained from 1H nmr studies. However, the cis:trans ratio of the Ala-Pro bond was not affected by N-methylation of the penultimate amino acid residue, which prevents formation of this hydrogen bond. Overall the experimental observations lead to the conclusion that the relative energies of the peptide conformations including cis or trans proline are mainly determined by intramolecular electrostatic interactions, whereas in the molecules considered, intramolecular hydrogen bonding is a consequence of specific peptide backbone conformations rather than a cause for the occurrence of energetically favored species. Independent support for this conclusion was obtained from model consideration which indicated that electrostatic interactions between the terminal carboxylic acid group and the carbonyl oxygen of the penultimate amino acid residue could indeed account for the observed relative conformational energies of the species containing cis and trans proline, respectively.  相似文献   

8.
An isolated uncharged hydrogen bond acceptor such as the carbonyl functionality of an aldehyde or a keto group is absent in natural amino acids. Although glutamine and asparagine are known to hydrogen bond through the amide carbonyl group in their side chains, they also possess the amide ? NH2 group, which can act as a hydrogen bond donor. This makes the structural study of peptides containing an oxo residue, with an isolated carbonyl group in the side chain, interesting. Here, we report the synthesis of δ‐ and ε‐oxo amino acids and their incorporation into oligopeptides as the N‐terminal residue. The resultant oxo peptides were extensively studied using X‐ray crystallography to understand the interactions offered by the oxo group in peptide crystals. We find that the oxo groups are capable of providing additional hydrogen bonding opportunities to the peptides, resulting in increased intermolecular interactions in crystals. The study thus offers avenues for the utilization of oxo residues to introduce intermolecular interactions in synthetic peptides.  相似文献   

9.
M Mutter  H Mutter  R Uhmann  E Bayer 《Biopolymers》1976,15(5):917-927
The conformation of polyethylene glycol-bound peptides, synthesized by the liquid-phase method, was investigated. This marcromolecular C-terminal protecting group is transparent in the visible and the ultraviolet range to 190 nm and solubilizes peptides in many different solvents. The CD spectra of the polymer-bound myoglobin sequence 66–73 and of the biologically active undecapeptide “substance P” were measured in each step of the synthesis. In both examples the formation of a secondary structure during the growth of the peptide chain was found. In the hydrophobic octapeptide containing the myoglobin sequence 66–73, the influence of either the blocked or the free N-terminal amino group on the conformation was observed. The blocked octapeptide in trifluoroethanol showed a higher degree of α-helix contribution than in its free state. The conformation of the polyethylene glycol-bound nona- and decaalanine in trifluoroethanol and water was determined. The peptide with a free amino end group has β-conformation in trifluoroethanol as well as in water. The corresponding N-Boc-protected derivatives show helical structure. The amino end group has a decisive influence on the formation of β-structure. The method of CD investigation of polymer-bound peptide sequences during the peptide synthesis in solution enables one to determine the influence of protecting groups and the chain end of a peptide on its conformation. It is also possible to study the relationship between the secondary structure, the chain length, and the kinetic of the coupling reaction in different solvents. Since the crystallization method for the liquid-phase peptide synthesis allows one to synthesize peptides in very short time, a new method of studying peptide conformations is opened.  相似文献   

10.
Use of the 4‐pyridylmethyl ester group for side‐chain protection of glutamic acid residues in solid‐phase peptide synthesis enables switching of the charge state of a peptide from negative to positive, thus making detection by positive ion mode ESI‐MS possible. The pyridylmethyl ester moiety is readily removed from peptides in high yield by hydrogenation. Combining the 4‐pyridylmethyl ester protecting group with benzyl ester protection reduces the number of the former needed to produce a net positive charge and allows for purification by RP HPLC. This protecting group is useful in the synthesis of highly acidic peptide sequences, which are often beset by problems with purification by standard RP HPLC and characterization by ESI‐MS. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
Short cyclic peptides have a great interest in therapeutic, diagnostic and affinity chromatography applications. The screening of ‘one‐bead‐one‐peptide’ combinatorial libraries combined with mass spectrometry (MS) is an excellent tool to find peptides with affinity for any target protein. The fragmentation patterns of cyclic peptides are quite more complex than those of their linear counterparts, and the elucidation of the resulting tandem mass spectra is rather more difficult. Here, we propose a simple protocol for combinatorial cyclic libraries synthesis and ring opening before MS analysis. In this strategy, 4‐hydroxymethylbenzoic acid, which forms a benzyl ester with the first amino acid, was used as the linker. A glycolamidic ester group was incorporated after the combinatorial positions by adding glycolic acid. The library synthesis protocol consisted in the following: (i) incorporation of Fmoc‐Asp[2‐phenylisopropyl (OPp)]‐OH to Ala‐Gly‐oxymethylbenzamide‐ChemMatrix, (ii) synthesis of the combinatorial library, (iii) assembly of a glycolic acid, (iv) couple of an Ala residue in the N‐terminal, (v) removal of OPp, (vi) peptide cyclisation through side chain Asp and N‐Ala amino terminus and (vii) removal of side chain protecting groups. In order to simultaneously open the ring and release each peptide, benzyl and glycolamidic esters were cleaved with ammonia. Peptide sequences could be deduced from the tandem mass spectra of each single bead evaluated. The strategy herein proposed is suitable for the preparation of one‐bead‐one‐cyclic depsipeptide libraries that can be easily open for its sequencing by matrix‐assisted laser desorption/ionisation MS. It employs techniques and reagents frequently used in a broad range of laboratories without special expertise in organic synthesis. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.

Non-proteinogenic amino acids are widely explored group of compounds due to their chemical properties and great potential of application in the combinatorial chemistry, medicinal investigation etc. Therefore the synthetic methods of their incorporation to the peptide chain are required. l-Mimosine, (S)-α-amino-β-(3-hydoxy-4-oxo-1,4-dihydropyridin-1-yl)-propanoic acid), is a plant amino acid, known to induce apoptosis in human pancreatic cancer xenografts. Here we present our investigations on the synthesis of mimosine-containing peptide and their ESI-MS/MS analysis. We successfully applied Fmoc-protected mimosine a with a free hydroxy ketone group for efficient peptide synthesis in the presence of HATU as a coupling reagent without the formation of side products. Additionally the tandem mass spectrometry analysis revealed the characteristic loss of the heterocyclic ring from mimosine residue side chain. The described method allows insertion of mimosine residue at any endo-position within a peptide sequence. The obtained results may be useful in the synthesis and mass spectrometry analysis of various mimosine-containing peptides.

  相似文献   

13.
DKP formation is a serious side reaction during the solid‐phase synthesis of peptide acids containing either Pro or Gly at the C‐terminus. This side reaction not only leads to a lower overall yield, but also to the presence in the reaction crude of several deletion peptides lacking the first amino acids. For the preparation of protected peptides using the Fmoc/tBu strategy, the use of a ClTrt‐Cl‐resin with a limited incorporation of the C‐terminal amino acid is the method of choice. The use of resins with higher loading levels leads to more impure peptide crudes. The use of HPLC‐ESMS is a useful method for analysing complex samples, such as those formed when C‐terminal Pro peptides are prepared by non‐optimized solid‐phase strategies. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
P Novak  I K Dev 《Journal of bacteriology》1988,170(11):5067-5075
The degradation of the prolipoprotein signal peptide in vitro by membranes, cytoplasmic fraction, and two purified major signal peptide peptidases from Escherichia coli was followed by reverse-phase liquid chromatography (RPLC). The cytoplasmic fraction hydrolyzed the signal peptide completely into amino acids. In contrast, many peptide fragments accumulated as final products during the cleavage by a membrane fraction. Most of the peptides were similar to the peptides formed during the cleavage of the signal peptide by the purified membrane-bound signal peptide peptidase, protease IV. Peptide fragments generated during the cleavage of the signal peptide by protease IV and a cytoplasmic enzyme, oligopeptidase A, were identified from their amino acid compositions, their retention times during RPLC, and knowledge of the amino acid sequence of the signal peptide. Both enzymes were endopeptidases, as neither dipeptides nor free amino acids were formed during the cleavage reactions. Protease IV cleaved the signal peptide predominantly in the hydrophobic segment (residues 7 to 14). Protease IV required substrates with hydrophobic amino acids at the primary and the adjacent substrate-binding sites, with a minimum of three amino acids on either side of the scissile bond. Oligopeptidase A cleaved peptides (minimally five residues) that had either alanine or glycine at the P'1 (primary binding site) or at the P1 (preceding P'1) site of the substrate. These results support the hypothesis that protease IV is the major signal peptide peptidase in membranes that initiates the degradation of the signal peptide by making endoproteolytic cuts; oligopeptidase A and other cytoplasmic enzymes further degrade the partially degraded portions of the signal peptide that may be diffused or transported back into the cytoplasm from the membranes.  相似文献   

15.
The unique physicochemical properties of a thioamide bond, which is an ideal isostere of an amide bond, have not been fully exploited because of the tedious synthesis of thionated amino acid building blocks. Here, we report a purification‐free and highly efficient synthesis of thiobenzotriazolides of Fmoc‐protected and orthogonally protected 20 naturally occurring amino acids including asparagine, glutamine, and histidine. The near‐quantitative conversion to the respective thioamidated peptides on solid support demonstrates the robustness of the synthetic route. Furthermore, the unaltered incorporation efficiency of thiobenzotriazolides from their stock solution till 48 h suggests their compatibility toward automated peptide synthesis. Finally, utilizing an optimized cocktail of 2% DBU + 5% piperazine for fast Fmoc‐deprotection, we report the synthesis of a thioamidated Pin1 WW domain and thioamidated GB1 directly on solid support.  相似文献   

16.
Halophilic proteins are characterized by high net negative charges and relatively small fraction of hydrophobic amino acids, rendering them aggregation resistant. These properties are also shared by histidine‐rich metal binding protein (HP) from moderate halophile, Chromohalobacter salexigens, used in this study. Here, we examined how halophilic proteins form amyloid fibrils in vitro. His‐tagged HP, incubated at pH 2.0 and 58°C, readily formed amyloid fibrils, as observed by thioflavin fluorescence, CD spectra, and transmission or atomic force microscopies. Under these low‐pH harsh conditions, however, His‐HP was promptly hydrolyzed to smaller peptides most likely responsible for rapid formation of amyloid fibril. Three major acid‐hydrolyzed peptides were isolated from fibrils and turned out to readily form fibrils. The synthetic peptides predicted to form fibrils in these peptide sequences by Waltz software also formed fibrils. Amyloid fibril was also readily formed from full‐length His‐HP when incubated with 10–20% 2,2,2‐trifluoroethanol at pH 7.8 and 25°C without peptide bond cleavage.  相似文献   

17.
TL Joseph  DP Lane  CS Verma 《PloS one》2012,7(8):e43985
Atomistic simulations of a set of stapled alpha helical peptides derived from the BH3 helix of MCL-1 (Stewart et al. (2010) Nat Chem Biol 6: 595-601) complexed to a fragment (residues 172-320) of MCL-1 revealed that the highest affinity is achieved when the staples engage the surface of MCL-1 as has also been demonstrated for p53-MDM2 (Joseph et al. (2010) Cell Cycle 9: 4560-4568; Baek et al. (2012) J Am Chem Soc 134: 103-106). Affinity is also modulated by the ability of the staples to pre-organize the peptides as helices. Molecular dynamics simulations of these stapled BH3 peptides were carried out followed by determination of the energies of interactions using MM/GBSA methods. These show that the location of the staple is a key determinant of a good binding stapled peptide from a bad binder. The good binder derives binding affinity from interactions between the hydrophobic staple and a hydrophobic patch on MCL-1. The position of the staple was varied, guiding the design of new stapled peptides with higher affinities.  相似文献   

18.
For every n amino acids linked in a protein there are n − 1 peptide bonds. The free energy of peptide bond hydrolysis and formation in aqueous solution defines the equilibrium position between peptide and amino acid hydrolysis products. Yet few experimental values exist. With a minimum of assumptions, this paper deduces the free energies of hydrolysis of a variety of peptide bonds. Formation of a dipeptide from two amino acids is about eight times more difficult than subsequent condensations of an amino acid to a dipeptide or longer chain. Condensation of an amino acid to a peptide of any size is five times more difficult than joining two smaller peptides of at least dipeptide size. Thus in an abiogenesis scenario there is a kind of nucleation in peptide bond formation with the initial condensation of two amino acids to yield a dipeptide more difficult than subsequent condensations to a growing chain. © 1998 John Wiley & Sons, Inc. Biopoly 45: 351–353, 1998  相似文献   

19.
This report describes a Dde resin based attachment strategy for inverse solid-phase peptide synthesis (ISPPS). This attachment strategy can be used for the synthesis of amino terminated peptides with side chains and the carboxyl terminus either protected or deprotected. Amino acid t-butyl esters were attached through their free amino group to the Dde resin. The t-butyl carboxyl protecting group was removed by 50% TFA, and inverse peptide synthesis cycles performed using an HATU/TMP based coupling method. Protected peptides were cleaved from the resin with dilute hydrazine. Side chain protecting groups could then be removed by treatment with TFMSA/TFA. The potential of this approach was demonstrated by the synthesis of several short protected and unprotected peptides in good yield and with low epimerization. Its potential for peptide mimetic synthesis was demonstrated by the synthesis of two peptide trifluoromethylketones.  相似文献   

20.
Aminoacyl-tRNAs are the biologically active substrates for peptide bond formation in protein synthesis. The stability of the acyl linkage in each aminoacyl-tRNA, formed through an ester bond that connects the amino acid carboxyl group with the tRNA terminal 3′-OH group, is thus important. While the ester linkage is the same for all aminoacyl-tRNAs, the stability of each is not well characterized, thus limiting insight into the fundamental process of peptide bond formation. Here, we show, by analysis of the half-lives of 12 of the 22 natural aminoacyl-tRNAs used in peptide bond formation, that the stability of the acyl linkage is effectively determined only by the chemical nature of the amino acid side chain. Even the chirality of the side chain exhibits little influence. Proline confers the lowest stability to the linkage, while isoleucine and valine confer the highest, whereas the nucleotide sequence in the tRNA provides negligible contribution to the stability. We find that, among the variables tested, the protein translation factor EF-Tu is the only one that can protect a weak acyl linkage from hydrolysis. These results suggest that each amino acid plays an active role in determining its own stability in the acyl linkage to tRNA, but that EF-Tu overrides this individuality and protects the acyl linkage stability for protein synthesis on the ribosome.  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