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 0 毫秒
1.
The tripeptide Bz-Arg-Gly-Asp(-OMe)-OH was synthesized by enzymatic method. Bz-Arg-Gly-OEt was synthesized by trypsin in ethanol containing 0.1 M Tris/HCl buffer (pH 8.0), and then H-Asp(-OMe)(2) was incorporated into the Bz-Arg-Gly-OEt using chymopapain in 0.25M CHES/NaOH buffer (pH = 9.0, EDTA 10 mM). The yield of Bz-Arg-Gly-OEt and Bz-Arg-Gly-Asp(-OMe)-OH were 80% and 70% using 1M Bz-Arg-OEt and 0.5M Bz-Arg-Gly-OEt, respectively. For Bz-Arg-Gly-OEt synthesis reaction at high concentrations of the substrates, the buffer content in ethanol was a key factor to determine the optimal reaction condition. In Bz-Arg-Gly-Asp(-OMe)-OH synthesis reaction, the yield was low in organic solvent due to various side products such as Bz-Arg-OH, Bz-Arg-Gly-OH, and Bz-Arg-Gly-Asp(-OMe)-Asp(-OMe)-OH, suggesting that chymopapain has a very broad substrate specificity of the S(1) site. The Bz-Arg-Gly-Asp(-OMe)-OH synthesis rate and its yield were dramatically elevated and the side reactions were reduced using only the CHES/NaOH buffer (pH = 9.0, EDTA 10 mM) as a reaction media. The final product Bz-Arg-Gly-Asp(-OMe)-OH was identified to be formed via C-terminal hydrolysis of Bz-Arg-Gly-Asp(-OMe)(2) after the nucleophile, H-Asp(-OMe)(2), was added.  相似文献   

2.
A new route was employed to synthesize RGD. First, Gly-Asp dipeptide was synthesized by a novel chemical method in two steps, including chloroacetylation of L-aspartic acid and ammonolysis of chloroacetyl L-aspartic acid. Second, Nalpha-Z- L-Arginine was reacted with Gly-Asp to synthesize RGD by the N-carboxyanhydride method. Less protected amino acids were used in this synthesis. This method possessed advantages of low cost, simplicity, and rapidity with a reasonable yield of 62% calculated from arginine. In addition, compared with the above method, a conventional solid phase method was also used to synthesize RGD, the yield was 75% calculated from the first amino acid anchored to resin.  相似文献   

3.
The tripeptide BzArgGlyAsp(NH(2))(2) was synthesized by a combination of chemical and enzymatic methods in this study. First of all, GlyAsp(NH(2))(2) was synthesized by a novel chemical method in three steps including chloroacetylation of L-aspartic acid, esterification of chloroacetyl L-aspartic acid and ammonolysis of chloroacetyl L-aspartic acid diethyl ester. Secondly, kinetically controlled synthesis of BzArgGlyAsp(NH(2))(2) catalyzed by trypsin in organic solvent was conducted. The optimum conditions are pH 8.0, 30 degrees C in ethanol/Tris-HCl buffer system (85:15, v/v) for 80 min in the maximum yield of 74.4%.  相似文献   

4.
Chen JZ  Han XW  Xie XQ 《Life sciences》2005,76(18):2053-2069
Anandamide (arachidonyl-ethanolamide, AEA) is an important endogenous cannabinoid ligand isolated from porcine brain. AEA has a flexible molecular structure with a series of four non-conjugated double bonds, a hydrophobic alkyl chain, and a carboxyamide head group. It is known that AEA binds to cannabinoid receptor and induces cannabimimetic activity. However, questions still remain about the three-dimensional arrangement of the pharmacophoric groups of AEA that facilitate its interaction with cannabinoid receptor, a member of transmembrane G-protein coupled receptors (GPCRs). Such information is of critical importance for the design of novel analogs of potential therapeutic values. In the present studies, we developed a combined approach of 2D high-resolution NMR and computer modeling to investigate conformational features of AEA in solution. The developed method and experimental data is then applied to study the structural properties of AEA in a membrane-like environment that will be reported elsewhere. In addition to the measured NOEs, the dihedral angle constraints were for the first time being used as experimentally-determined structural constraints for performing molecular dynamics simulations to refine the NMR-determined AEA conformations. Our results showed that AEA prefers an extended pseudo-helical conformation in solution with two oxygen atoms pointing towards the same side and a straight pentyl chain, which was an averaged conformation observed on the basis of NMR time scale. The results were correlated to the computer predicted AEA models reported by others. The established NMR-based computational approach provides an alternative way to explore further the detailed conformational properties of AEA that encodes important pharmacophoric and conformational information regarding the activation of cannabinoid receptors.  相似文献   

5.
The tripeptide Bz-Arg-Gly-Asp(NH2)OH was synthesized by a combination of chemical and enzymatic methods in this study. Firstly, Gly-Asp-(NH2)2 was synthesized by a novel chemical method in three steps including chloroacetylation of l-aspartic acid, esterification of chloroacetyl l-aspartic acid and ammonolysis of chloroacetyl l-aspartic acid diethyl ester. Secondly, the linkage of the third amino acid (Bz-Arg-OEt) to Gly-Asp-(NH2)2 was completed by enzymatic method under kinetic control condition. An industrial alkaline protease alcalase was used in water–organic cosolvents systems. The synthesis reaction conditions were optimized by examining the effects of several factors including water content, temperature, pH and reaction time on the yield of the synthesis product Bz-Arg-Gly-Asp(NH2)OH. The optimum conditions are pH 8.0, 35 °C, in ethanol/Tris–HCl buffer system (85:15, v/v), 8 h with the tripeptide yield of 73.6%.  相似文献   

6.
Arg-Gly-Asp (RGD) peptides contain an aspartic acid residue that is highly susceptible to chemical degradation and leads to the loss of biological activity. Our hypothesis is that cyclization of RGD peptides via disulphide bond linkage can induce structural rigidity, thereby preventing degradation mediated by the aspartic acid residue. In this paper, we compared the solution stability of a linear peptide (Arg-Gly-Asp-Phe-OH; 1) and a cyclic peptide (cyclo-(1, 6)-Ac-Cys-Arg-Gly-Asp-Phe-Pen-NH2; 2) as a function of pH and buffer concentration. The decomposition of both peptides was studied in buffers ranging from pH 2-12 at 50 degrees C. Reversed-phase HPLC was used as the main tool in determining the degradation rates and pathways of both peptides. Fast atom bombardment mass spectrometry (FAB-MS), electrospray ionization mass spectrometry (ESI-MS), matrix-assisted laser desorption/ionization-time of flight (MALDI-TOF) mass spectrometry, liquid chromatography-mass spectrometry (LC-MS), and one- and two-dimensional nuclear magnetic resonance spectroscopy (NMR) were used to characterize peptides 1 and 2 and their degradation products. In addition, co-elution with authentic samples was used to identify degradation products. Both peptides displayed pseudo-first-order kinetics at all pH values studied. The cyclic peptide 2 appeared to be 30-fold more stable than the linear peptide 1 at pH 7. The degradation mechanisms of linear (1) and cyclic (2) peptides primarily involved the aspartic acid residue. However, above pH 8 the stability of the cyclic peptide decreased dramatically due to disulphide bond degradation. Both peptides also exhibited a change in degradation mechanism upon an increase in pH. The increase in stability of cyclic peptide 2 compared to linear peptide 1, especially at neutral pH, may be due to decreased structural flexibility imposed by the ring. This rigidity would prevent the Asp side chain carboxylic acid from orientating itself in the appropriate position for attack on the peptide backbone.  相似文献   

7.
Possible conformations of penicillin G; d and l isomers of ampicillin; α-amino-α-methyl-benzyl penicillins and 3- pyridyl methyl penicillin have been studied by an energy minimization procedure using empirical potential functions. The preferred conformations of these antibiotics have been correlated with their biological activity. The conformational requirement of the antibiotic to be active against Gram-positive and Gram-negative (β-lactamase-negative) bacterial strains seems to be the same. The reduced activity of penicillin G against Gram-negative bacteria has been attributed to its lower ability to permeate the outer membrane. The flexibility of the sidechains of these antibiotics is also shown to be important for the desired biological activity.  相似文献   

8.
9.
Six retro-inverso tri- and tetrapeptide analogues of RGD were prepared and their anti-aggregatory activity was determined by platelet aggregation tests in comparison with the corresponding parent peptides. An efficient method for the introduction of a malonyl-aspartic residue into a peptide chain is described for the first time. A 2-3-fold decrease in potency or total loss of bioactivity was observed with the new peptides; structure-activity relationships are discussed.  相似文献   

10.
11.
The lone-pair orbital interactions arising in a phosphodiester are incorporated into semiempirical conformational energy calculations using a unifold “torsional potential” around the virtual bond linking the ester oxygen atoms. The results explain the observed experimental data better than other methods.  相似文献   

12.
Park HS  Kim C  Kang YK 《Biopolymers》2002,63(5):298-313
The conformational study on Arg-Gly-Asp (RGD)-containing tetrapeptides in the unhydrated and hydrated states has been carried out using the force field ECEPP/3 and the hydration shell model. The tetrapeptides studied here are H-RGDX-OH (X = Trp, Tyr, Phe, Leu, Val, Cys, Gln, and Ser), which show the inhibitory activity for binding of fibrinogen to platelets in the order of RGDW approximately equal to RGDY approximately equal to RGDF approximately equal to RGDL > RGDV > or = RGDC > or = RGDQ > or = RGDS. The backbone conformations with two C(7) backbone-to-backbone hydrogen bonds between Asp and Arg residues and between Xaa and Gly residues are in common most probable for the RGD sequence of RGDX tetrapeptides in the hydrated state. The dominant beta-turns for RGDX are found to be the types V' and IV at Gly-Asp and Asp-Xaa sequences, respectively, which are quite similar to the types II' and I (or II), respectively. However, it cannot be ruled out that the extended conformations are also remarkably feasible for RGDX tetrapeptides in water by peering the distributions of backbone conformations. These calculated results are consistent with the experimental results on RGD-containing proteins and conformationally constrained RGD-containing peptides. The reason why the RGDX becomes more potent as the side chain of the X residue is more hydrophobic may be ascribed to that the more hydrophobic is the residue X, the more populated are beta-turn structures for the Gly-Asp sequence. The hydrophobic side chain of X residue exposed to water is likely to interact with the hydrophobic region of receptor easily.  相似文献   

13.
Bluetongue virus (BTV) is an arthropod-borne virus transmitted by Culicoides species to vertebrate hosts. The double-capsid virion is infectious for Culicoides vector and mammalian cells, while the inner core is infectious for only Culicoides-derived cells. The recently determined crystal structure of the BTV core has revealed an accessible RGD motif between amino acids 168 to 170 of the outer core protein VP7, whose structure and position would be consistent with a role in cell entry. To delineate the biological role of the RGD sequence within VP7, we have introduced point mutations in the RGD tripeptide and generated three recombinant baculoviruses, each expressing a mutant derivative of VP7 (VP7-AGD, VP7-ADL, and VP7-AGQ). Each expressed mutant protein was purified, and the oligomeric nature and secondary structure of each was compared with those of the wild-type (wt) VP7 molecule. Each mutant VP7 protein was used to generate empty core-like particles (CLPs) and were shown to be biochemically and morphologically identical to those of wt CLPs. However, when mutant CLPs were used in an in vitro cell binding assay, each showed reduced binding to Culicoides cells compared to wt CLPs. Twelve monoclonal antibodies (MAbs) was generated using purified VP7 or CLPs as a source of antigen and were utilized for epitope mapping with available chimeric VP7 molecules and the RGD mutants. Several MAbs bound to the RGD motif on the core, as shown by immunogold labeling and cryoelectron microscopy. RGD-specific MAb H1.5, but not those directed to other regions of the core, inhibited the binding activity of CLPs to the Culicoides cell surface. Together, these data indicate that the RGD motif present on BTV VP7 is responsible for Culicoides cell binding activity.  相似文献   

14.
15.
The conformational study on cyclic Ac-Cys-Pro-Xaa-Cys-NHMe (Ac-CPXC-NHMe; X=Ala, Val, Leu, Aib, Gly, His, Phe, Tyr, Asn and Ser) peptides has been carried out using the Empirical Conformational Energy Program for Peptides, version 3 (ECEPP/3) force field and the hydration shell model in the unhydrated and hydrated states. This work has been undertaken to investigate structural implications of the CPXC sequence as the chain reversal for the initiation of protein folding and as the motif for active site of disulfide oxidoreductases. The backbone conformation DAAA is commonly the most feasible for cyclic CPXC peptides in the hydrated state, which has a type I beta-turn at the Pro-Xaa sequence. The proline residue and the hydrogen bond between backbones of two cystines as well as the formation of disulfide bond appear to play a role in stabilizing this preferred conformation of cyclic CPXC peptides. However, the distributions of backbone conformations and beta-turns may indicate that the cyclic CPXC peptide seems to exist as an ensemble of beta-turns and coiled conformations in aqueous solution. The intrinsic stability of the cyclic CPXC motif itself for the active conformation seems to play a role in determining electrochemical properties of disulfide oxidoreductases.  相似文献   

16.
The objective of this study was to evaluate the relationship between conformational flexibility and solution stability of a linear RGD peptide (Arg-Gly-Asp-Phe-OH; 1) and a cyclic RGD peptide (cyclo-(1, 6)-Ac-Cys-Arg-Gly-Asp-Phe-Pen-NH2; 2); as a function of pH. Previously, it was found that cyclic peptide 2 was 30-fold more stable than linear peptide 1. Therefore, this study was performed to explain the increase in chemical stability based on the preferred conformation of the peptides. Molecular dynamics simulations and energy minimizations were conducted to evaluate the backbone flexibility of both peptides under simulated pH conditions of 3, 7 and 10 in the presence of water. The reactive sites for degradation for both molecules were also followed during the simulations. The backbone of linear peptide 1 exhibited more flexibility than that of cyclic peptide 2, which was reflected in the rotation about the phi and psi dihedral angles. This was further supported by the low r.m.s. deviations of the backbone atoms for peptide 2 compared with those of peptide 1 that were observed among structures sampled during the molecular dynamics simulations. The presence of a salt bridge between the side chain groups of the Arg and Asp residues was also indicated for the cyclic peptide under simulated conditions of neutral pH. The increase in stability of the cyclic peptide 2 compared with the linear peptide 1, especially at neutral pH, is due to decreased structural flexibility imposed by the ring, as well as salt bridge formation between the side chains of the Arg and Asp residues in cyclic peptide 2. This rigidity would prevent the Asp side chain carboxylic acid from orienting itself in the appropriate position for attack on the peptide backbone.  相似文献   

17.
The preferred conformations of the nonpolar β and γ (hydrocarbon) chains in phospholipids have been derived using EHT and CNDO calculations. These calculations indicate that the possible conformations of phospholipids are highly restricted. The calculations find support from X-ray diffraction studies and NMR measurements on model compounds. When considering conformations relevant to structures in cell membranes, a further selection is possible because of the fact that in aqueous solutions hydrophobic interactions stabilize an arrangement where the hydrocarbon chains (β and γ) are stacked almost parallel to one another, leading to a bilayer structure. The various models for β and γ-chains which satisfy this condition have been compared and it has been shown that of these only four are favoured by energy considerations. These arrangements differ from one another in the orientation of the β-chain and γ-chains in the interior of the bilayer structure. A low energy pathway connects these conformations and thus the molecule can easily flip from one stable bilayer arrangement to another. The possible conformations of the polar group (α) are likewise restricted. The proposed model provides explanations to a number of dynamic and static properties of phospholipids, in particular to the observed NMR coupling constants, 1H and 13C relaxation times, studies based on ESR spin labels and the observed X-ray diffraction results on model compounds.  相似文献   

18.
Important aspects in detailed nmr analyses of the conformations of linear peptides are discussed using enkephalin and the α-mating factor of Saccharomyces cerevisiae as examples. The cationic, dipolar, and anionic forms in dimethyl sulfoxide solution may be identified by ir analyses. Because of the electrostatic interaction between the N- and C-terminal groups, the dipolar form of enkephalin takes the folded conformation, as well as extended conformation(s), in dimethyl sulfoxide solution. Such conformational equilibrium is responsible for anomalous temperature dependences and solvent-composition dependences of the amide and Cα proton chemical shifts. Active analogs, enkephalinamide and enkephalinol, take extended conformation(s) in solution. These opioid peptides probably take a specific active conformation upon binding with a receptor. For the α-mating factor and active peptide analogs in aqueous solution, a folded conformation with two βturn structures is responsible for the biological activity.  相似文献   

19.
Mimetics of the RGD tripeptide are described that are potent, selective antagonists of the integrin receptor, alpha(v)beta(3). The use of the 5,6,7,8-tetrahydro[1,8]naphthyridine group as a potency-enhancing N-terminus is demonstrated. Two 3-substituted-3-amino-propionic acids previously contained in alpha(IIb)beta(3) antagonists were utilized to enhance binding affinity and functional activity for the targeted receptor. Further affinity increases were then achieved through the use of cyclic glycyl amide bond constraints.  相似文献   

20.
A conformational analysis of protected glutamate homo-oligopeptides Z-[Glu(OEt)]n-OEt (n = 2–7) was carried out in chloroform solution using high-resolution 1H-nmr spectroscopy. At dilute peptide concentrations, the backbone NH and α-CH resonances are well resolved and can be assigned by combining extensive homonuclear decoupling experiments with data for co-oligopeptide derivatives. The structure of these peptides in solution was then assessed using information from chemical shifts, coupling constants, temperature coefficients, and titration of each oligomer with trifluoroacetic acid (TFA). The di- and tripeptides are found to be in disordered forms in deuterochloroform (CDCl3) and CDCl3/TFA mixtures. The tetrapeptide exhibits a folded structure with intramolecular hydrogen bonding at Glu2 in CDCl3 and undergoes a transition to increasingly disordered forms as TFA is added. The pentamer to heptamer show a folded structure with a strong intramolecular hydrogen bond at Glu2 and a weaker hydrogen bond at Glu3, which are disrupted as these peptides go to random coils at high TFA/CDCl3 ratios. In addition, the N-terminal portions of these glutamate peptides appear to be involved in side chain–main chain interactions. The results support the hypothesis that protected linear homo-oligopeptides may possess two or more segments of conformation with intramolecular folding preferred near the N-terminal portion.  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