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 15 毫秒
1.
The 4‐methoxybenzyloxymethyl (MBom) group was introduced at the Nπ‐position of the histidine (His) residue by using a regioselective procedure, and its utility was examined under standard conditions used for the conventional and the microwave (MW)‐assisted solid phase peptide synthesis (SPPS) with 9‐fluorenylmethyoxycarbonyl (Fmoc) chemistry. The Nπ‐MBom group fulfilling the requirements for the Fmoc strategy was found to prevent side‐chain‐induced racemization during incorporation of the His residue even in the case of MW‐assisted SPPS performed at a high temperature. In particular, the MBom group proved to be a suitable protecting group for the convergent synthesis because it remains attached to the imidazole ring during detachment of the protected His‐containing peptide segments from acid‐sensitive linkers by treatment with a weak acid such as 1% trifluoroacetic acid in dichloromethane. We also demonstrated the facile synthesis of Fmoc‐His(π‐MBom)‐OH with the aid of purification procedure by crystallization to effectively remove the undesired τ‐isomer without resorting to silica gel column chromatography. This means that the present synthetic procedure can be used for large‐scale production without any obstacles. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
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   

3.
An efficient method of peptide thioester synthesis is described. The reaction is based on an N‐4,5‐dimethoxy‐2‐mercaptobenzyl (Dmmb) auxiliary‐assisted NS acyl shift reaction after assembling a peptide chain by Fmoc‐solid phase peptide synthesis. The Dmmb‐assisted NS acyl shift reaction proceeded efficiently under mildly acidic conditions, and the peptide thioester was obtained by treating the resulting S‐peptide with sodium 2‐mercaptoethanesulfonate. No detectable epimerization of the amino acid residue adjacent to the thioester moiety in the case of Leu was found. The reactions were also amenable to the on‐resin preparation of peptide thioesters. The utility was demonstrated by the synthesis of a 41‐mer peptide thioester, a phosphorylated peptide thioester and a 33‐mer peptide thioester containing a trimethylated lysine residue. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
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   

5.
Two new rigid bi‐aromatic linkers for synthesis of peptide arrays by SPOT methodology were obtained from cellulose treated with 2,4‐dichloro‐6‐methoxy‐1,3,5‐triazine. Reaction with m‐phenylenediamine gave non‐cleavable TYPE I linker which enabled attachment of the peptides via resistant to harsh reaction conditions amide, ether, and amine bonds. Reaction with 3‐Fmoc‐aminobenzoic acid followed by thermal isomerization of the intermediate “superactive” ester producing an amide‐like bond gave TYPE II linker that was very stable during peptide synthesis. However, the peptide was cleavable, with fragment of the linker, in the presence of 1 M LiOH solution. The uniform loading of the cellulose and efficient synthesis of the peptide array was achieved by using N‐(4,6‐dimethoxy‐1,3,5‐triazin‐1‐yl)‐N‐methylmorpholinium 4‐toluenesulfonate as the coupling reagent.  相似文献   

6.
In our efforts to develop a universal solution to the problem of aspartimide formation in Fmoc SPPS, we investigated the application of our new β‐trialkylmethyl protected aspartic acid building blocks to the synthesis of peptides containing the Asp‐Gly motif. The Nα‐Fmoc aspartic acid β‐tri‐(ethyl/propyl/butyl)methyl esters were used in the synthesis of the classic model peptide scorpion toxin II (VKDGYI), and their effectiveness in minimising aspartimide formation during extended piperidine treatments was evaluated. Furthermore, we compared their efficacy against that of the commonly used approach of adding acids to the Fmoc deprotection solution. Finally, we applied our aspartic acid building blocks to the stepwise Fmoc SPPS of teduglutide, a human GLP‐2 analogue, whose synthesis is made challenging by extensive aspartimide formation. In all experiments, our approach led to almost complete reduction of aspartimide formation with accompanied suppression of aspartic acid epimerisation. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

7.
In plants, 3‐deoxy‐d ‐manno‐oct‐2‐ulosonic acid (Kdo) is a monosaccharide that is only found in the cell wall pectin, rhamnogalacturonan‐II (RG‐II). Incubation of 4‐day‐old light‐grown Arabidopsis seedlings or tobacco BY‐2 cells with 8‐azido 8‐deoxy Kdo (Kdo‐N3) followed by coupling to an alkyne‐containing fluorescent probe resulted in the specific in muro labelling of RG‐II through a copper‐catalysed azide–alkyne cycloaddition reaction. CMP‐Kdo synthetase inhibition and competition assays showing that Kdo and D‐Ara, a precursor of Kdo, but not L‐Ara, inhibit incorporation of Kdo‐N3 demonstrated that incorporation of Kdo‐N3 occurs in RG‐II through the endogenous biosynthetic machinery of the cell. Co‐localisation of Kdo‐N3 labelling with the cellulose‐binding dye calcofluor white demonstrated that RG‐II exists throughout the primary cell wall. Additionally, after incubating plants with Kdo‐N3 and an alkynated derivative of L‐fucose that incorporates into rhamnogalacturonan I, co‐localised fluorescence was observed in the cell wall in the elongation zone of the root. Finally, pulse labelling experiments demonstrated that metabolic click‐mediated labelling with Kdo‐N3 provides an efficient method to study the synthesis and redistribution of RG‐II during root growth.  相似文献   

8.
1. The extent of racemization and the coupling yield in peptide synthesis were studied under high dilution conditions. The azide method yielded the best results. 2. Five linear penta-peptide precursors related to gramicidin S were subjected to cyclization in order to study how the difference in the sequence influences the yield and the ratio of cyclic dimer to monomer. The azide with the sequence of -L -Pro-L -Val-L -Orn(Z)-L -Leu-D -Phe- afforded diZ-gramicidin S in a high yield of 63%. 3. Alternaria mali toxin III, a cyclotetradepsipeptide phytotoxin, was synthesized. The activated linear tetradepsipeptide containing a D -Dap(Z) (N3-Z-D -2,3-diaminopropionic acid) residue at the N-terminus afforded the cyclic precursor (53%). The Dap residue in the precursor was converted into a ΔAla residue by Hofmann degradation to give the desired product.  相似文献   

9.
Human glycodelin consists of 162 amino acid residues and two N‐linked glycans at Asn28 and Asn63. In this study, we synthesized it by a fully convergent strategy using native chemical ligation (NCL) in N to C direction. The four peptide segments corresponding to 1–31, 32–65, 66–105 and 106–162 sequences were synthesized by 9‐fluorenylmethoxycarbonyl based solid‐phase peptide synthesis. At the C‐terminus of the second segment, N‐ethyl‐S‐acetamidomethyl‐cysteine was attached as a post‐ligation thioesterification device. The N‐terminal two segments were condensed by the homocysteine‐mediated NCL at Leu‐Met site, and the product was methylated to convert homocysteine to methionine. After deprotection of acetamidomethyl group on the N‐ethylcysteine residue, the peptide was thioesterified by N‐alkylcysteine‐assisted method. The product was then ligated with the C‐terminal half, which was obtained by the NCL of third and fourth segments, to give the full‐length glycodelin. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
Native chemical ligation is widely used for the convergent synthesis of proteins. The peptide thioesters required for this process can be challenging to produce, particularly when using Fmoc‐based solid‐phase peptide synthesis. We have previously reported a route to peptide thioesters, following Fmoc solid‐phase peptide synthesis, via an NS acyl shift that is initiated by the presence of a C‐terminal cysteine residue, under mildly acidic conditions. Under typical reaction conditions, we occasionally observed significant thioester hydrolysis as a consequence of long reaction times (~48 h) and sought to accelerate the reaction. Here, we present a faster route to peptide thioesters, by replacing the C‐terminal cysteine residue with selenocysteine and initiating thioester formation via an NSe acyl shift. This modification allows thioester formation to take place at lower temperatures and on shorter time scales. We also demonstrate how application of this strategy also accelerates peptide cyclization, when a linear precursor is furnished with an N‐terminal cysteine and C‐terminal selenocysteine. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
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   

12.
A peptide, N‐Ac‐Phe‐Tyr‐NH2, with angiotensin I‐converting enzyme (ACE) inhibitor activity was synthesized by an α‐chymotrypsin‐catalyzed condensation reaction of N‐acetyl phenylalanine ethyl ester (N‐Ac‐Phe‐OEt) and tyrosinamide (Tyr‐NH2). Three kinds of solvents: a Tris–HCl buffer (80 mM, pH 9.0), dimethylsulfoxide (DMSO), and acetonitrile were employed in this study. The optimum reaction solvent component was determined by simplex centroid mixture design. The synthesis efficiency was enhanced in an organic‐aqueous solvent (Tris‐HCl buffer: DMSO: acetonitrile = 2:1:1) in which 73.55% of the yield of N‐Ac‐Phe‐Tyr‐NH2 could be achieved. Furthermore, the effect of reaction parameters on the yield was evaluated by response surface methodology (RSM) using a central composite rotatable design (CCRD). Based on a ridge max analysis, the optimum condition for this peptide synthesis included a reaction time of 7.4 min, a reaction temperature of 28.1°C, an enzyme activity of 98.9 U, and a substrate molar ratio (Phe:Tyr) of 1:2.8. The predicted and the actual (experimental) yields were 87.6 and 85.5%, respectively. The experimental design and RSM performed well in the optimization of synthesis of N‐Ac‐Phe‐Tyr‐NH2, so it is expected to be an effective method for obtaining a good yield of enzymatic peptide. © 2012 American Institute of Chemical Engineers Biotechnol. Prog., 2012  相似文献   

13.
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   

14.
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   

15.
The structure of N‐nitroso, N‐(2‐chloroethyl), N′,N′‐dibenzylsulfamid (CENS) was established by X‐ray crystallography. The atomic coordinates, factors of isotropic thermal agitation, bond lengths and valence angles were determined. The solvent effects on the electronic absorption and fluorescence spectra of CENS were investigated at room temperature. The effects of solvent polarity and of hydrogen bonding were interpreted by means of linear solvation energy relationships (LSERs). Multiple linear regression analysis indicated that the hydrogen donation properties of the solvent play an important role in determining the position of the absorption maximum, while the classical polarity of the medium is the only dominating parameter in determining the emission maximum and the Stokes' shift. Complexation of the investigated compound by two different transition metal ions was studied. Fluorescence measurements show that fluorescence quenching by cobalt(II) is more important than that by copper(II). This phenomenon can be attributed to good stereo‐structural matching between the electronic configuration of the Co2+ ion and the active site distribution of CENS in aqueous solution. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

16.
A toxin with four disulfide bridges from Tityus serrulatus venom was able to compete with 125I‐kaliotoxin on rat brain synaptosomal preparations, with an IC50 of 46 nM . The obtained amino acid sequence and molecular mass are identical to the previously described butantoxin. Enzymatic cleavages in the native peptide followed by mass spectrometry peptide mapping analysis were used to determine the disulfide bridge pattern of α‐KTx12?1. Also, after the cleavage of the first six N‐terminal residues, including the unusual disulfide bridge which forms an N‐terminus ring, the potency of the cleaved peptide was found to decrease about 100 fold compared with the native protein. Copyright © 2003 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
Crustacean insulin‐like androgenic gland factor (IAG) of Macrobrachium rosenbergii, a heterodimeric peptide having both four disulfide bonds and an N‐linked glycan, was synthesized by the combination of solid‐phase peptide synthesis and the regioselective disulfide formation reactions. The disulfide isomer of IAG could also be synthesized by the same manner. The conformational analysis of these peptides by circular dichroism (CD) spectral measurement indicated that the disulfide bond arrangement affected the peptide conformation in IAG. On the other hand, the N‐linked glycan attached at A chain showed no effect on CD spectra of IAG. This is the first report for the chemical synthesis of insulin‐like heterodimeric glycopeptide having three interchain disulfides, and the synthetic strategy shown here might be useful for the synthesis of other glycosylated four‐disulfide insulin‐like peptides.  相似文献   

18.
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   

19.
To photomodulate the interaction of the phosphatidylinositol 3‐kinase SH3 domain with a peptide ligand, a cyclic peptide (cyclic‐1) with a photolabile side chain‐to‐side chain linker was synthesized. The conformation of cyclic‐1 differs from that of the parent linear peptide, but becomes identical by UV‐irradiation. Accordingly, the binding affinity of cyclic‐1 to the SH3 domain increased upon conversion of the cyclic to a linear flexible structure by irradiation (Kd: 3.4 ± 1.7 and 0.9 ± 0.3 mM , respectively). These results confirm the usefulness of a photocleavable peptide for photocontrol of peptide–protein interactions. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
Aggregation, orientation and dynamics of transmembrane helices are of relevance for protein function and transmembrane signaling. To explore the interactions of transmembrane helices and the interdependence of peptide structure and lipid composition of the membranes, β‐peptides were explored as model transmembrane domains. Various hydrophobic β‐peptide sequences were synthesized by solid phase peptide synthesis. Conformational analyses of β‐peptide helices were performed in organic solvents (methanol and 2,2,2‐trifluoroethanol) and in large unilamellar liposomes (dimyristoylphosphatidylcholine, dipalmitoylphosphatidylcholine and dioleoylphosphatidylcholine) indicating 12‐ and 14‐helix conformations, depending on β3‐amino acid sequences. The intrinsic tryptophan fluorescence of β3‐homotryptophan units inserted in the center or near the end of the sequence was used to verify the membrane insertion of the β‐peptides. A characteristic blue shift with peripheral β3‐homotryptophan compared with β‐peptides with central tryptophan served as indication for a transmembrane orientation of the β‐peptides within the lipid bilayer. Copyright © 2016 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