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.
Site‐specific labeling of synthetic peptides carrying N‐methoxyglycine (MeOGly) by isothiocyanate is demonstrated. A nonapeptide having MeOGly at its N‐terminus was synthesized by the solid‐phase method and reacted with phenylisothiocyanate under various conditions. In acidic solution, the reaction specifically gave a peptide having phenylthiourea structure at its N‐terminus, leaving side chain amino group intact. The synthetic human β‐defensin‐2 carrying MeOGly at its N‐terminus or the side chain amino group of Lys10 reacted with phenylisothiocyanate or fluorescein isothiocyanate also at the N‐methoxyamino group under the same conditions, demonstrating that this method is generally useful for the site‐specific labeling of linear synthetic peptides as well as disulfide‐containing peptides. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
The Pictet–Spengler (PS) cyclizations of β3hTrp derivatives as arylethylamine substrates were performed with L‐α‐amino and D‐α‐amino aldehydes as carbonyl components. During the PS reaction, a new stereogenic center was created, and the mixture of cis/trans 1,3‐disubstituted 1,2,3,4‐tetrahydro‐β‐carbolines was obtained. The ratio of cis/trans diastereomers depends on the stereogenic centre of used amino aldehyde and the size of substituents. It was confirmed by 1H and 2D NMR (ROESY) spectra. The conformations of cyclic products were studied by 2D NMR ROESY spectra. Products of the PS condensation after removal of protecting group(s) can be incorporated into a peptide chain as tryptophan mimetics with the possibility of the β‐turn induction. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

3.
During the final step of t‐Boc/Bzl, solid‐phase peptide synthesis (SPPS)‐protecting groups from amino acids (aa) side chains must be removed from the target peptides during cleavage from the solid support . These reaction steps involve hydrolysis with hydrogen fluoride (HF) in the presence of a nucleophile (scavenger), whose function is to trap the carbocations produced during SN1‐type reactions. Five peptide sequences were synthesised for evaluating p‐methoxyphenol effectiveness as a potent scavenger. After the synthesis, the resin–peptide was then separated into two equal parts to be cleaved using two scavengers: conventional reactive p‐cresol (reported in the literature as an effective acyl ion eliminator) and p‐methoxyphenol (hypothesised as fulfilling the same functions as the routinely used scavenger). Detailed analysis of the electrostatic potential map (EPM) revealed similarities between these two nucleophiles, regarding net atomic charge, electron density distribution, and similar pKa values. Good scavenger efficacy was observed by chromatography and mass spectrometry results for the synthesised molecules, which revealed that p‐methoxyphenol can be used as a potent scavenger during SPPS by t‐Boc/Bzl strategy, as similar results were obtained using the conventional scavenger.  相似文献   

4.
The increasing interest in click chemistry and its use to stabilize turn structures led us to compare the propensity for β‐turn stabilization of different analogs designed as mimics of the β‐turn structure found in tendamistat. The β‐turn conformation of linear β‐amino acid‐containing peptides and triazole‐cyclized analogs were compared to ‘conventional’ lactam‐ and disulfide‐bridged hexapeptide analogs. Their 3D structures and their propensity to fold in β‐turns in solution, and for those not structured in solution in the presence of α‐amylase, were analyzed by NMR spectroscopy and by restrained molecular dynamics with energy minimization. The linear tetrapeptide Ac‐Ser‐Trp‐Arg‐Tyr‐NH2 and both the amide bond‐cyclized, c[Pro‐Ser‐Trp‐Arg‐Tyr‐D ‐Ala] and the disulfide‐bridged, Ac‐c[Cys‐Ser‐Trp‐Arg‐Tyr‐Cys]‐NH2 hexapeptides adopt dominantly in solution a β‐turn conformation closely related to the one observed in tendamistat. On the contrary, the β‐amino acid‐containing peptides such as Ac‐(R)‐β3‐hSer‐(S)‐Trp‐(S)‐β3‐hArg‐(S)‐β3‐hTyr‐NH2, and the triazole cyclic peptide, c[Lys‐Ser‐Trp‐Arg‐Tyr‐βtA]‐NH2, both specifically designed to mimic this β‐turn, do not adopt stable structures in solution and do not show any characteristics of β‐turn conformation. However, these unstructured peptides specifically interact in the active site of α‐amylase, as shown by TrNOESY and saturation transfer difference NMR experiments performed in the presence of the enzyme, and are displaced by acarbose, a specific α‐amylase inhibitor. Thus, in contrast to amide‐cyclized or disulfide‐bridged hexapeptides, β‐amino acid‐containing peptides and click‐cyclized peptides may not be regarded as β‐turn stabilizers, but can be considered as potential β‐turn inducers. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

5.
This work reports an efficient Lewis acid catalysed N‐methylation procedure of lipophilic α‐amino acid methyl esters in solution phase. The developed methodology involves the use of the reagent system AlCl3/diazomethane as methylating agent and α‐amino acid methyl esters protected on the amino function with the (9H‐fluoren‐9‐yl)methanesulfonyl (Fms) group. The removal of Fms protecting group is achieved under the same conditions to those used for Fmoc removal. Thus the Fms group can be interchangeable with the Fmoc group in the synthesis of N‐methylated peptides using standard Fmoc‐based strategies. Finally, the absence of racemization during the methylation reaction and the removal of Fms group were demonstrated by synthesising a pair of diastereomeric dipeptides. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

6.
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   

7.
Glyceroacetonide–Oxyma [(2,2‐dimethyl‐1,3‐dioxolan‐4‐yl)methyl 2‐cyano‐2‐(hydroxyimino)acetate ( 1 )] displayed remarkable physico‐chemical properties as an additive for peptide‐forming reactions. Although racemization‐free amide‐forming reactions have been established for N‐urethane‐protected α‐amino acids with EDCI, 1 , and NaHCO3 in water or DMF‐water media, amide‐forming reactions of N‐acyl‐protected α‐amino acids and segment couplings of oligopeptides still require further development. Diethylphosphoryl–glyceroacetonide–oxyma (DPGOx 3 ) exhibits relative stability in aprotic solvents and is an effective coupling reagent for N‐acyl‐protected α‐amino acids and oligo peptide segments. The conditions reported here is also effective in lactam‐forming reactions. Unlike most of the reported coupling reagents, simple aqueous work‐up procedures can remove the reagents and by‐products generated in the reactions. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
The measurement of glycated hemoglobin A1c (HbA1c) has important implications for diagnosis of diabetes and assessment of treatment effectiveness. We proposed specific sequence motifs to identify enzymes that oxidize glycated compounds from genome database searches. The gene encoding a putative fructosyl amino acid oxidase was found in the Phaeosphaeria nodorum SN15 genome and successfully expressed in Escherichia coli. The recombinant protein (XP_001798711) was confirmed to be a novel fructosyl peptide oxidase (FPOX) with high specificity for α‐glycated compounds, such as HbA1c model compounds fructosyl‐αN‐valine (f‐αVal) and fructosyl‐αN‐valyl‐histidine (f‐αVal‐His). Unlike previously reported FPOXs, the P. nodorum FPOX has a Km value for f‐αVal‐His (0.185 mM) that is considerably lower than that for f‐αVal (0.458 mM). Based on amino acid sequence alignment, three dimensional structural modeling, and site‐directed mutagenesis, Gly60 was found to be a determining residue for the activity towards f‐αVal‐His. A flexible surface loop region was also found to likely play an important role in accepting f‐αVal‐His. Biotechnol. Bioeng. 2010; 106: 358–366. © 2010 Wiley Periodicals, Inc.  相似文献   

9.
A procedure for the synthesis of a11C‐labeled oligopeptide containing [1‐11C]1,2,3,4‐tetrahydro‐β‐carboline‐3‐carboxylic acid ([1‐11C]Tpi) from the corresponding Trp?HCl‐containing peptides has been developed involving a Pictet‐Spengler reaction with [11C]formaldehyde. The synthesis of [1‐11C]Tpi from Trp and [11C]formaldehyde was examined as a model reaction with the aim of developing a facile and effective method for the labeling of peptides with carbon‐11. The Pictet‐Spengler reaction of Trp and [11C]formaldehyde in acidic media (TsOH or HCl) afforded the desired [1‐11C]Tpi in a moderate radiochemical yield. Herein, the application of a Pictet‐Spengler reaction to an aqueous solution of Trp?HCl gave the desired product with a radiochemical yield of 45.2%. The RGD peptide cyclo[Arg‐Gly‐Asp‐D‐Tyr‐Lys] was then selected as a substrate for the labeling reaction with [11C]formaldehyde. The radiolabeling of a Trp?HCl‐containing RGD peptide using the Pictet‐Spengler reaction was successful. Furthermore, the remote‐controlled synthesis of a [1‐11C]Tpi‐containing RGD peptide was attempted by using an automatic production system to generate [11C]CH3I. The radiochemical yield of the [1‐11C]Tpi‐containing RGD at the end of synthesis (EOS) was 5.9 ± 1.9% (n = 4), for a total synthesis time of about 35 min. The specific activity was 85.7 ± 9.4 GBq/µmol at the EOS. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
N‐[1‐(4‐(4‐fluorophenyl)‐2,6‐dioxocyclohexylidene)ethyl] (Fde) protected amino acids have been prepared and applied in solid‐phase peptide synthesis monitored by gel‐phase 19F NMR spectroscopy. The Fde protective group could be cleaved with 2% hydrazine or 5% hydroxylamine solution in DMF as determined with gel‐phase 19F NMR spectroscopy. The dipeptide Ac‐L ‐Val‐L ‐Val‐NH2 12 was constructed using Fde‐L ‐Val‐OH and no noticeable racemization took place during the amino acid coupling with N,N′‐diisopropylcarbodiimide and 1‐hydroxy‐7‐azabenzotriazole or Fde deblocking. To extend the scope of Fde protection, the hydrophobic nonapeptide LLLLTVLTV from the signal sequence of mucin MUC1 was successfully prepared using Fde‐L ‐Leu‐OH at diagnostic positions. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
N‐terminal modification of peptides by unnatural amino acids significantly affects their enzymatic stability, conformational properties and biological activity. Application of N‐amidino‐amino acids, positively charged under physiological conditions, can change peptide conformation and its affinity to the corresponding receptor. In this article, we describe synthesis of short peptides, containing a new building block—N‐amidino‐pyroglutamic acid. Although direct guanidinylation of pyroglutamic acid and oxidation of N‐amidino‐proline using RuO4 did not produce positive results, N‐amidino‐Glp‐Phe‐OH was synthesized on Wang polymer by cyclization of α‐guanidinoglutaric acid residue. In the course of synthesis, it was found that literature procedure of selective Boc deprotection using TMSOTf/TEA reagent is accompanied by concomitant side reaction of triethylamine alkylation by polymer linker fragment. It should be mentioned that independently from cyclization time and coupling agent (DIC or HCTU), the lactam formation was incomplete. Separation of the cyclic product from the linear precursor was achieved by HPLC in ammonium formate buffer at pH 6. HPLC analysis showed N‐amidino‐Glp‐Phe‐OH stability at acidic and physiological pH and fast ring opening in water solution at pH 9. The suggested method of N‐amidino‐Glp residue formation can be applied in the case of short peptide chains, whereas synthesis of longer ones will require fragment condensation approach. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
Via a Mannich reaction involving a dibenzyliminium species and the titanium enolates of Evans' chiral acylated oxazolidinones the β2‐amino acids (R)‐ and (S)‐Fmoc‐β2homovaline and (R)‐Fmoc‐β2homoleucine are synthesized. These building blocks were used, in combination with commercially available α‐ and β3‐amino acids, for the synthesis of the cyclo‐(αβ3αβ2α)2 peptide 2 and the cyclo‐(αβ2αβ3α)2 peptides 3 – 5 . The peptides 2 – 5 were screened for their ability to inhibit a small panel of Gram‐negative and Gram‐positive bacterial strains.  相似文献   

13.
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   

14.
Tropoelastin is the primary component of elastin, which forms the elastic fibers that make up connective tissues. The hydrophobic domains of tropoelastin are thought to mediate the self‐assembly of elastin into fibers, and the temperature‐mediated self‐assembly (coacervation) of one such repetitive peptide sequence (VPGVG) has been utilized in various bio‐applications. To elucidate a mechanism for coacervation activity enhancement and to develop more potent coacervatable elastin‐derived peptides, we synthesized two series of peptide analogs containing an aromatic amino acid, Trp or Tyr, in addition to Phe‐containing analogs and tested their functional characteristics. Thus, position 1 of the hydrophobic pentapeptide repeat of elastin (X1P2G3V4G5) was substituted by Trp or Tyr. Eventually, we acquired a novel, short Trp‐containing elastin‐derived peptide analog (WPGVG)3 with potent coacervation ability. From the results obtained during this process, we determined the importance of aromaticity and hydrophobicity for the coacervation potency of elastin‐derived peptide analogs. Generally, however, the production of long‐chain synthetic polypeptides in quantities sufficient for commercial use remain cost‐prohibitive. Therefore, the identification of (WPGVG)3, which is a 15‐mer short peptide consisting simply of five natural amino acids and shows temperature‐dependent self‐assembly activity, might serve as a foundation for the development of various kinds of biomaterials. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
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   

16.
The huntingtin protein is characterized by a segment of consecutive glutamines (QN) that is responsible for its fibrillation. As with other amyloid proteins, misfolding of huntingtin is related to Huntington's disease through pathways that can involve interactions with phospholipid membranes. Experimental results suggest that the N‐terminal 17‐amino‐acid sequence (httNT) positioned just before the QN region is important for the binding of huntingtin to membranes. Through all‐atom explicit solvent molecular dynamics simulations, we unveil the structure and dynamics of the httNTQN fragment on a phospholipid membrane at the atomic level. We observe that the insertion dynamics of this peptide can be described by four main steps—approach, reorganization, anchoring, and insertion—that are very diverse at the atomic level. On the membrane, the httNT peptide forms a stable α‐helix essentially parallel to the membrane with its nonpolar side‐chains—mainly Leu‐4, Leu‐7, Phe‐11 and Leu‐14—positioned in the hydrophobic core of the membrane. Salt‐bridges involving Glu‐5, Glu‐12, Lys‐6, and Lys‐15, as well as hydrogen bonds involving Thr‐3 and Ser‐13 with the phospholipids also stabilize the structure and orientation of the httNT peptide. These observations do not significantly change upon adding the QN region whose role is rather to provide, through its hydrogen bonds with the phospholipids' head group, a stable scaffold facilitating the partitioning of the httNT region in the membrane. Moreover, by staying accessible to the solvent, the amyloidogenic QN region could also play a key role for the oligomerization of httNTQN on phospholipid membranes. Proteins 2014; 82:1409–1427. © 2014 Wiley Periodicals, Inc.  相似文献   

17.
The ON acyl transfer reaction has gained significant popularity in peptide and medicinal chemistry. This reaction has been successfully applied to the synthesis of difficult sequence‐containing peptides, cyclic peptides, epimerization‐free fragment coupling and more recently, to switchable peptide polymers. Herein, we describe a related strategy to facilitate the synthesis and purification of a hydrophobic stapled peptide. The staple consists of a serine linked through an amide bond formed from its carboxylic acid function and the side chain amino group of diaminopropionic acid and through an ester bond formed from its amino group and the side chain carboxylic acid function of aspartic acid. The α‐amino group of serine was protonated during purification. Interestingly, when the peptide was placed at physiological pH, the free amino group initiated the O‐N shift reducing the staple length by one atom, leading to a more hydrophobic stapled peptide. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
Three fluorescein derivatives of human insulin (HI, 1 ) labeled at positions NαA1, NαB1 and NεB29 respectively, were synthesized using an N‐trifluoroacetyl‐based protecting group scheme. The Tfa protecting group introduced by reaction with ethyl trifluoroacetate was found to be stable in aqueous and organic media and efficiently removed under mild basic conditions. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
Lactoferricin (LfB) is a 25‐residue innate immunity peptide released by pepsin from the N‐terminal region of bovine lactoferrin. A smaller amidated peptide, LfB6 (RRWQWR‐NH2) retains antimicrobial activity and is thought to constitute the “antimicrobial active‐site” (Tomita, Acta Paediatr Jpn. 1994; 36 : 585–91). Here we report on N‐acylation of 1‐Me‐Trp5‐LfB6, Cn‐RRWQ[1‐Me‐W]R‐NH2, where Cn is an acyl chain having n = 0, 2, 4, 6 or 12 carbons. Tryptophan 5 (Trp5) was methylated to enhance membrane binding and to allow for selective deuteration at that position. Peptide/lipid interactions of Cn‐RRWQ[1‐Me‐W ]R‐NH2 (deuterated 1‐Me‐Trp5 underlined), were monitored by solid state 31P NMR and 2H NMR. The samples consisted of macroscopically oriented bilayers of mixed neutral (dimyristoylphosphatidylcholine, DMPC) and anionic (dimyristoylphosphatidylglycerol, DMPG) lipids in a 3:1 ratio with Cn‐RRWQ[&1‐Me‐W ]R‐NH2 peptides added at a 1:25 peptide to lipid ratio. 2H‐NMR spectra reveal that the acylated peptides are well aligned in DMPC:DMPG bilayers. The 2H NMR quadrupolar splittings suggest that the 1‐Me‐Trp is located in a motionally restricted environment, indicating partial alignment at the membrane interface. 31P‐NMR spectra reveal that the lipids are predominantly in a bilayer configuration, with little perturbation by the peptides. Methylation alone, in C0‐RRWQ[1‐Me‐W ]R‐NH2, resulted in a 3–4 fold increase in antimicrobial activity against E. coli. N‐acylation with a C12 fatty acid enhanced activity almost 90 fold. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
Sunflower trypsin inhibitor‐1 (SFTI‐1), a bicyclic tetradecapeptide, has become a versatile tool as a scaffold for the development of the inhibitors of therapeutically relevant serine proteases, among them matriptase and kallikreins. Herein, we report the rational design of potent monocyclic and bicyclic inhibitors of human matriptase‐1. We found that the presence of positive charge and lack of bulky residues at the peptide N‐terminus is required for the maintenance of inhibitory activity. Replacement of the N‐terminal glycine residue by lysine allowed for the chemical conjugation with a fluorophor via the ε‐amino group without significant loss of inhibitory activity. Head‐to‐tail and side‐chain‐to‐tail cyclization resulted in potent inhibitors with comparable activities against matriptase‐1. The most potent synthetic bicyclic inhibitor found in this study (Ki = 2.6 nM at pH 7.6) is a truncated version of SFTI‐1 (cyclo‐KRCTKSIPPRCH) lacking a C‐terminal proline and aspartate residue. It combines an internal disulfide bond with a peptide macrocycle that is formed through side‐chain‐to‐tail cyclization of the ε‐amino group of an N‐terminal lysine and a C‐terminal proline. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  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