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 46 毫秒
1.
Aza‐glycinyl dipeptides are useful building blocks for the synthesis of a diverse array of azapeptides. The construction of the aza‐glycine residue is however challenging, because of the potential for side reactions, such as those leading to formation of oxadiazalone, hydantoin and symmetric urea by‐products. Employing N,N′‐disuccinimidyl carbonate to activate benzophenone hydrazone, we have developed a more efficient approach for the synthesis of aza‐glycinyl dipeptides. Alkylation of the semicarbazone of the resulting protected aza‐glycinyl dipeptides using tetraethylammonium hydroxide and propargyl bromide provided an efficient entry into the aza‐propargylglycinyl peptide building blocks, which have served previously in various reactions including Sonogashira cross‐couplings, dipolar cycloadditions and intramolecular exo‐dig cycloadditions to furnish a variety of azapeptide building blocks. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
The present work is devoted to the synthesis, conformational analysis, and stereodynamic study of aza‐β3‐cyclodipeptides. This pseudopeptidic ring shows E/Z hydrazide bond isomerism, eight‐membered ring conformation, and chirotopic nitrogen atoms, all of which are elements that are prone to modulate the ring shape. The (E,E) twist boat conformation observed in the solid state by X‐ray diffraction is also the ground conformation in solution, and emerges as the lowest in energy when using quantum chemical calculations. The relative configuration associated with ring chirality and with the two nitrogen chiral centers is governed by steric crowding and adopts the (P)SNSN/(M)RNRN combination which projects side chains in equatorial position. The nitrogen pyramidal inversion (NPI) at the two chiral centers is correlated with the ring reversal. The process is significantly hindered as was shown by VT‐NMR experiments run in C2D2Cl4, which did not make it possible to determine the barrier to inversion. Finally, these findings make it conceivable to resolve enantiomers of aza‐β3‐cyclodipeptides by modulating the backbone decoration appropriately. Chirality 25:341–349, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

3.
N‐(tert‐butyloxycarbonyl) or N‐(9‐fluorenylmethoxycarbonyl) dipeptides with C‐terminal (Z)‐α,β‐didehydrophenylalanine (?ZPhe), (Z)‐α,β‐didehydrotyrosine (?ZTyr), (Z)‐α,β‐didehydrotryptophan (?ZTrp), (Z)‐α,β‐didehydromethionine (?ZMet), (Z)‐α,β‐didehydroleucine (?ZLeu), and (Z/E)‐α,β‐didehydroisoleucine (?Z/EIle) were synthesised from their saturated analogues via oxidation of intermediate 2,5‐disubstituted‐oxazol‐5‐(4H)‐ones (also known as azlactones) with pyridinium tribromide followed by opening of the produced unsaturated oxazol‐5‐(4H)‐one derivatives in organic‐aqueous solution with a catalytic amount of trifluoroacetic acid or by a basic hydrolysis. In all cases, a very strong preference for Z isomers of α,β‐didehydro‐α‐amino acid residues was observed except of the ΔIle, which was obtained as the equimolar mixture of Z and E isomers. Reasons for the (Z)‐stereoselectivity and the increased stability of the aromatic α,β‐didehydro‐α‐amino acid residue oxazol‐5‐(4H)‐ones over the corresponding aliphatic ones are also discussed. It is the first use of such a procedure to synthesise peptides with the C‐terminal unsaturated residues and a peptide with 2 consecutive ΔPhe residues. This approach is very effective especially in the synthesis of peptides with aliphatic α,β‐didehydro‐α‐amino acid residues that are difficult to obtain by other methods. It allowed the first synthesis of the ?Met residue. It is also more cost‐effective and less laborious than other synthesis protocols. The dipeptide building blocks obtained were used in the solid‐phase synthesis of model peptides on a polystyrene‐based solid support. Peptides containing aromatic α,β‐didehydro‐α‐amino acid residues were obtained with PyBOP or TBTU as a coupling agent with good yields and purities. In the case of aliphatic α,β‐didehydro‐α‐amino acid residues, a good efficiency was achieved only with DPPA as a coupling agent.  相似文献   

4.
Synthetic peptides reproducing the helix‐loop‐helix (HLH) domains of the Id proteins fold into highly stable helix bundles upon self‐association. Recently, we have shown that the replacement of the dipeptide Val‐Ser at the loop–helix‐2 junction with the corresponding O‐acyl iso‐dipeptide leads to a completely unfolded state that only refolds after intramolecular ON acyl migration. Herein, we report on an Id HLH analog based on the substitution of the Pro‐Ser motif at the helix‐1–loop junction with the corresponding O‐acyl iso‐dipeptide. This analog has been successfully synthesized by solid‐phase Fmoc chemistry upon suppression of DKP formation. No secondary structure could be detected for the O‐acyl iso‐peptide before its conversion into the native form by ON acyl shift. These results show that the loop–helix junctions are determinant for the folded/unfolded state of the Id HLH domain. Further, despite the high risk of DKP formation, peptides containing O‐acyl iso‐Pro‐Ser/Thr units are synthetically accessible by Fmoc chemistry. Copyright © 2010 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.
《Chirality》2017,29(8):458-463
Cinchona alkaloid derivatives as organocatalysts were applied in the asymmetric aza‐ Friedel–Crafts reaction of α‐naphthol with aryl aldimines. The desired chiral aminoarylnaphthols were obtained in 85% enantiomeric excess.  相似文献   

7.
The importance of dipeptides both in medicinal and pharmacological fields is well documented and many efforts have been made to find simple and efficient methods for their synthesis. For this reason, we have investigated the synthesis of α‐N‐protected dipeptide acids by reacting the easily accessible mixed anhydride of α‐N‐protected amino acids with free amino acids under different reaction conditions. The combination of TBA‐OH and DMSO has been found to be the best to overcome the low solubility of amino acids in organic solvents. Under these experimental conditions, the homogeneous phase condensation reaction occurs rapidly and without detectable epimerization. The present method is also applicable to side‐chain unprotected Tyr, Trp, Glu, and Asp but not Lys. This latter residue is able to engage two molecules of mixed anhydride giving the corresponding isotripeptide. Moreover, the applicability of this protocol for the synthesis of tri‐ and tetrapeptides has been tested. This approach reduces the need for protecting groups, is cost effective, scalable, and yields dipeptide acids that can be used as building blocks in the synthesis of larger peptides. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
A series of terminally blocked peptides (to the pentamer level) from l ‐Ala and the cyclic Cα,α‐disubstituted Gly residue Afc and one Gly/Afc dipeptide have been synthesized by solution method and fully characterized. The molecular structure of the amino acid derivative Boc‐Afc‐OMe and the dipeptide Boc‐Afc‐Gly‐OMe were determined in the crystal state by X‐ray diffraction. In addition, the preferred conformation of all of the model peptides was assessed in deuterochloroform solution by FT‐IR absorption and 1H‐NMR. The experimental data favour the conclusion that the Afc residue tends to adopt either the fully‐extended (C5) or a folded/helical structure. In particular, the former conformation is highly populated in solution and is also that found in the crystal state in the two compounds investigated. A comparison with the structural propensities of the strictly related Cα,α‐disubstituted Gly residues Ac5c and Dϕg is made and the implications for the use of the Afc residue in conformationally constrained analogues of bioactive peptides are briefly examined. A spectroscopic (UV absorption, fluorescence, CD) characterization of this novel aromatic Cα,α‐disubstituted Gly residue is also reported. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
Young Kee Kang  In Kee Yoo 《Biopolymers》2014,101(11):1077-1087
Conformational preferences of 9‐ and 14‐helix foldamers have been studied for γ‐dipeptides of 2‐aminocyclohexylacetic acid (γAc6a) residues such as Ac‐(γAc6a)2‐NHMe ( 1 ), Ac‐(Cα‐Et‐γAc6a)2‐NHMe ( 2 ), Ac‐(γAc6a)2‐NHBn ( 3 ), and Ac‐(Cα‐Et‐γAc6a)2‐NHBn ( 4 ) at the M06‐2X/cc‐pVTZ//M06‐2X/6‐31 + G(d) level of theory to explore the influence of substituents on their conformational preferences. In the gas phase, the 9‐helix foldamer H9 and 14‐helix foldamer H14‐z are found to be most preferred for dipeptides 2 and 4 , respectively, as for dipeptides 1 and 3 , which indicates no remarkable influence of the Cα‐ethyl substitution on conformational preferences. The benzyl substitution at the C‐terminal end lead H14‐z to be the most preferred conformer for dipeptides 3 and 4 , whereas it is H9 for dipeptides 1 and 2 , which can be ascribed to the favored C? H···π interactions between the cyclohexyl group of the first residue and the C‐terminal benzyl group. There are only marginal changes in backbone structures and the distances and angles of H‐bonds for all local minima by Cα‐ethyl and/or benzyl substitutions. Although vibrational frequencies and intensities of the dipeptide 4 calculated at both M06‐2X/6‐31 + G(d) and M05‐2X/6‐31 + G(d) levels of theory are consistent with observed results in the gas phase, H14‐z is predicted to be most preferred by ΔG only at the former level of theory. Hydration did not bring the significant changes in backbone structures of helix foldamers for both dipeptide 1 and 4 . It is expected that the different substitutions at the C‐terminal end lead to the different helix foldamers, which may increase the resistance of helical structures to proteolysis and provide the more surface to the helical structures suitable for molecular recognition. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 1077–1087, 2014.  相似文献   

10.
The protein tyrosine phosphatase SHP‐1 plays an important role in many physiological and pathophysiological processes. This phosphatase is activated through binding of ligands to its SH2‐domains, mainly to the N‐terminal one. Based on a theoretical docking model, backbone‐to‐side chain cyclized octapeptides were designed as ligands. Assembly of such modelled structures required the synthesis of N‐functionalized tyrosine derivatives and their incorporation into the sequence. Because of difficulties encountered in the condensation of N‐protected amino acids to the N‐alkylated tyrosine‐peptide we synthesized and used preformed dipeptide building units. As all attempts to obtain phosphorylated dipeptide units failed, the syntheses had to be performed with a free phenolic function. Use of different N‐alkyl or cycloalkyl residues in the N‐functionalized side chains allowed to investigate the effect of ring size, flexibility and hydrophobicity of formed lactam bridges on stimulatory activity. All tested linear and cyclic octapeptides stimulate the phosphatase activity of SHP‐1. Stimulatory activities of cyclic ligands increase with the chain length of the lactam bridges resulting in increased flexibility and better entropic preformation of the binding conformation. The strong activity of some cyclic octapeptides supports the modelled structure. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
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   

12.
In the synthesis of dipeptides (Boc‐AA1‐AA2‐OPac: AA1 and AA2 represent amino acids) protected by phenacyl (Pac) ester, amines and solid bases as the base for the conversion of the trifluoroacetic acid (TFA) salt of the amino component (TFA·H‐AA2‐OPac) into the corresponding free amino component (H‐AA2‐OPac) were examined. The synthesis of a dipeptide (Boc‐Ala‐Gly‐OPac) using amines for the conversion afforded an unsatisfactory yield with by‐products. On the other hand, the use of neutral alumina‐supported Na2CO3 (Na2CO3/n‐Al2O3) as a solid base for the conversion provided the dipeptide in a quantitative yield without by‐products. The application of Na2CO3/n‐Al2O3 to the synthesis of some dipeptides protected by Pac ester gave the desired peptides in excellent yields. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
Structural modification of the peptide backbone via N‐methylation is a powerful tool to modulate the pharmacokinetic profile and biological activity of peptides. Here we describe a rapid and highly efficient microwave(MW)‐assisted Fmoc/tBu solid‐phase method to prepare short chain N‐methyl‐rich peptides, using Rink amide p‐methylbenzhydrylamine (MBHA) resin as solid‐phase support. This method produces peptides in high yield and purity, and reduces the time required for Fmoc‐N‐methyl amino acid coupling. Copyright © 2010 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.
Thermal treatment of short‐chain oligopeptides is able to initiate the process of their self‐assembly with the formation of organic nanostructures with unique properties. On the other hand, heating can lead to a chemical reaction with the formation of new substances with specific properties and ability to form structures with different morphology. Therefore, in order to have a desired process, researcher needs to find its temperature range. In the present work, cyclization of L‐isoleucyl‐L‐alanine dipeptide in the solid state upon heating was studied. Kinetic parameters of this reaction were estimated within the approaches of the nonisothermal kinetics. The correlation between side chain structure of dipeptides and temperature of their cyclization in the solid state was found for the first time. This correlation may be used to predict the temperature, at which dipeptide self‐assembly changes to chemical reaction. The differences in self‐assembly of linear and cyclic dipeptides were demonstrated using atomic force microscopy. The effect of dipeptide concentration in a source solution and an organic solvent used on self‐assembly of dipeptides was shown. The new information obtained on the thermal properties and self‐assembly of linear and cyclic forms of L‐isoleucyl‐L‐alanine may be useful for the design of new nanomaterials based on oligopeptides, as well as for the synthesis of cyclic oligopeptides.  相似文献   

16.
A series of Fmoc‐Phe(4‐aza‐C60)‐OH of fullerene amino acid derived peptides have been prepared by solid phase peptide synthesis, in which the terminal amino acid, Phe(4‐aza‐C60)‐OH, is derived from the dipolar addition to C60 of the Fmoc‐Nα‐protected azido amino acids derived from phenylalanine: Fmoc‐Phe(4‐aza‐C60)‐Lys3‐OH ( 1 ), Fmoc‐Phe(4‐aza‐C60)‐Pro‐Hyp‐Lys‐OH ( 2 ), and Fmoc‐Phe(4‐aza‐C60)‐Hyp‐Hyp‐Lys‐OH ( 3 ). The inhibition constant of our fullerene aspartic protease PRIs utilized FRET‐based assay to evaluate the enzyme kinetics of HIV‐1 PR at various concentrations of inhibitors. Simulation of the docking of the peptide Fmoc‐Phe‐Pro‐Hyp‐Lys‐OH overestimated the inhibition, while the amino acid PRIs were well estimated. The experimental results show that C60‐based amino acids are a good base structure in the design of protease inhibitors and that their inhibition can be improved upon by the addition of designer peptide sequences. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
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   

18.
To prevent aspartimide formation and related side products in Asp‐Xaa, particularly Asp‐Gly‐containing peptides, usually the 2‐hydroxy‐4‐methoxybenzyl (Hmb) backbone amide protection is applied for peptide synthesis according to the Fmoc‐protocols. In the present study, the usefulness of the recently proposed acid‐labile dicyclopropylmethyl (Dcpm) protectant was analyzed. Despite the significant steric hindrance of this bulky group, N‐terminal H‐(Dcpm)Gly‐peptides are quantitatively acylated by potent acylating agents, and alternatively the dipeptide Fmoc‐Asp(OtBu)‐(Dcpm)Gly‐OH derivative can be used as a building block. In contrast to the Hmb group, Dcpm is inert toward acylations, but is readily removed in the acid deprotection and resin‐cleavage step. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
Reversed‐phase high‐pressure liquid chromatography analysis and purification of three hydrophobic, aggregation‐prone peptides, composed mainly of the transmembrane (TM) sequence, were performed using elution systems containing 2,2,2‐trifluoroethanol (TFE). The addition of 10–16% TFE to a common mobile phase, such as a water/acetonitrile/propanol (PrOH) or a water/PrOH/formic acid system, markedly improved the chromatographic separation of these peptides. The superior performance of TFE‐containing systems in separating peptides over water/PrOH/formic acid systems [Bollhagen R. et al., J. Chromatogr. A, 1995; 711 : 181–186.] clearly demonstrated that adding TFE to the mobile phase is one of best methods for TM‐peptide purification. Characterization of the potential side reactions using MALDI and ESI‐LIT/Orbitrap mass spectrometry indicated that prolonged incubation of peptides in a mixture of TFE–formic acid possibly induces O‐formylation of the Ser residue and N‐formylation of the N‐terminus of peptides. The conditions for selective removal of the formyl groups from TM peptides were also screened. We believe that these results will expand our ability to analyze and prepare hydrophobic, aggregation‐prone TM peptides and proteins. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
The use of lanthanides in preference to radioisotopes as probes for various biological assays has gained enormous popularity. The introduction of lanthanide chelates to peptides/proteins can be carried out either in solution using a commercially available labelling kit or by solid‐phase peptide synthesis using an appropriate lanthanide chelate. Herein, a detailed protocol for the latter is provided for the labelling of peptides or small proteins with diethylenetriamine‐N, N, N″, N″‐tetra‐tert‐butyl acetate‐N′‐acetic acid (DTPA) chelate or other similar chelates on a solid support using a chimeric insulin‐like peptide composed of human insulin‐like peptide 5 (INSL5) A‐chain and relaxin‐3 B‐chain as a model peptide. Copyright © 2011 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