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.
Continuing our research on Mycobacteria kansasii phagocytosis inhibition, we have examined in that context three series of peptides derived from the RGDVY and GRGD sequences. It was found that the levels of the inhibitory activity depend on the amino acid composition as well as on the particular peptide sequence. Distinct inhibitory activity was found in the case of thymopentin (RKDVY), the active fragment of thymopoietin. In this case the Mycobacterium phagocytosis inhibition should be combined with general immunostimulatory activity of RKDVY peptide. Our examination of a series of GRGDV analogs with a successively prolonged oligo-Gly linker inserted into the peptide chain showed that the distance between the Arg and Asp residues required for such an activity should be about 9A.  相似文献   

2.
Cleavage of specific peptide bonds occurs with aging in the alpha A subunit of bovine alpha-crystallin. One of the breaks occurs at residue Asn-101. This same residue undergoes in vivo deamidation, isomerization, and racemization. Deamidation and isomerization are known to occur via succinimide ring formation of labile asparagine residues. Model studies on peptides have shown that imide formation can also lead to peptide bond cleavage (Geiger, T., and Clarke, S. (1987) J. Biol. Chem. 262, 785-794). In that case, both asparagine and aspartic acid amide would be expected as C termini of the truncated polypeptide, and this is indeed the case in the alpha A-(1-101)-chain. This thus represents a first example of nonenzymatic in vivo peptide bond cleavage in an aging protein through the formation of a succinimide intermediate. In addition, we found that in bovine lens no detectable conversion (through the action of protein-carboxyl methyltransferase) of isoaspartyl to normal aspartyl residues occurs in vivo after deamidation of Asn-101.  相似文献   

3.
One mechanism for the spontaneous degradation of polypeptides is the intramolecular attack of the peptide bond nitrogen on the side chain carbonyl carbon atom of aspartic acid and asparagine residues. This reaction results in the formation of succinimide derivatives and has been shown to be largely responsible for the racemization, isomerization, and deamidation of these residues in several peptides under physiological conditions (Geiger, T. & Clarke, S. J. Biol. Chem. 262, 785-794 (1987]. To determine if similar reactions might occur in proteins, I examined the sequence and conformation about aspartic acid and asparagine residues in a sample of stable, well-characterized proteins. There did not appear to be any large bias against dipeptide sequences that readily form succinimides in small peptides. However, it was found that aspartyl and asparaginyl residues generally exist in native proteins in conformations where the peptide bond nitrogen atom cannot approach the side chain carbonyl carbon to form a succinimide ring. These orientations also represent energy minimum states, and it appears that this factor may account for a low rate of spontaneous damage to proteins by succinimide-linked reactions. The presence of aspartic acid and asparagine residues in other conformations, such as those in partially denatured, conformationally flexible regions, may lead to more rapid succinimide formation and contribute to the degradation of the molecule. The possible role of isoimide intermediates, formed by the attack of the peptide oxygen atom on the side chain carboxyl group, in protein racemization, isomerization, and deamidation is also considered.  相似文献   

4.
A novel cytotoxic linear peptide, IsCT, was characterized from scorpion Opisthacanthus madagascariensis. It is a linear peptide with a molecular weight of 1501.9 Da composed of 13 amino acid residues without cysteines. MS/MS analysis showed that its C-terminal is amidated. The identity of IsCT is re-confirmed by comparing the chemical synthesized peptide with the natural one. IsCT demonstrated antimicrobial activity against both gram-positive and gram-negative bacteria and hemolytic activity to sheep red blood cells. Also, it can release histamine from rat peritoneal mast cells. The CD absorption suggested that IsCT had an alpha-helix configuration in aqueous TFE. IsCT is one of the shortest natural cytotoxic peptides described, and it will be a suitable model for studying peptide-lipid interactions.  相似文献   

5.
We have isolated and purified adrenal chromogranin A (Ch A) for the purpose of making structural comparisons to parathyroid secretory protein-I (SP-I), because our earlier data indicated these two molecules may be the same protein. An improved purification step, using high-performance liquid chromatography (HPLC), has enabled us to demonstrate that both SP-I and Ch A consists of two species, one of approximately 72,000 Da and one of approximately 66,000 Da. The amino acid composition is the same for all four species. The difference in molecular mass is assumed to be due to carbohydrate content. Cyanogen bromide digestion of each of the four samples, followed by HPLC separation of the generated peptides, resulted in a chromatographic profile that was the same for each digest. Amino acid analysis of the eight peptide fragments obtained from each digest indicates that both species of Ch A and both species of SP-I yielded the same peptide mixtures following this cleavage reaction. One large (approximately 50,000 Da) CNBr peptide was obtained and seven smaller ones, one of which contains cysteine. The large fragment behaved similarly to the intact molecule in a radioimmunoassay. HPLC separation of tryptic digests of Ch A (72,000 Da) and SP-I (72,000 Da) also resulted in elution profiles that were very similar to each other. Amino acid analysis revealed 23 peptides common to each digest. Ch A contained four peptides ranging in size from 4 to 30 residues that were not observed in the SP-I digest. SP-I contained two peptides, each with about 30 residues, that were not found in the Ch A digest. Nothing unusual was noted in any of the uncommon peptides. Thus, both a chemical and an enzymatic digestion of these molecules followed by analysis of the peptides generated, indicates that SP-I and Ch A are nearly identical homologs.  相似文献   

6.
The structure of alpha-hemoglobin stabilizing protein (AHSP), a molecular chaperone for free alpha-hemoglobin, has been determined using NMR spectroscopy. The protein native state shows conformational heterogeneity attributable to the isomerization of the peptide bond preceding a conserved proline residue. The two equally populated cis and trans forms both adopt an elongated antiparallel three alpha-helix bundle fold but display major differences in the loop between the first two helices and at the C terminus of helix 3. Proline to alanine single point mutation of the residue Pro-30 prevents the cis/trans isomerization. The structure of the P30A mutant is similar to the structure of the trans form of AHSP in the loop 1 region. Both the wild-type AHSP and the P30A mutant bind to alpha-hemoglobin, and the wild-type conformational heterogeneity is quenched upon complex formation, suggesting that just one conformation is the active form. Changes in chemical shift observed upon complex formation identify a binding interface comprising the C terminus of helix 1, the loop 1, and the N terminus of helix 2, with the exposed residues Phe-47 and Tyr-51 being attractive targets for molecular recognition. The characteristics of this interface suggest that AHSP binds at the intradimer alpha1beta1 interface in tetrameric HbA.  相似文献   

7.
A simple and unambiguous method for the detection of the amino acids tyrosine and methionine in peptide structures has been developed. The procedure, which was applied in studies of opioid peptides, is based on continuous-flow fast atom bombardment mass spectrometry (CF-FAB-MS) following chemical modification of the residue to be analyzed. Thus, for the detection of tyrosine, modification reactions such as acetylation or non-radioactive iodination were performed prior to analysis by CF-FAB-MS. O-Acetylation of the tyrosine residue with N-acetylimidazole was accompanied by a shift of 42 Da in the molecular mass of the peptide under investigation. This modification was reversed by treatment with hydroxylamine hydrochloride. Incorporation of iodine resulted in a molecular weight shift of 126 Da per iodine atom. Methionine residues were detected in methionine-enkephalin-containing peptides following S-oxidation with hydrogen peroxide. The procedures described may have a wide application in peptide chemistry, particularly for the identification of peptide fragments containing the above residues, e.g. in studies of processing or degradation of the enkephalins or other neuropeptides (e.g. endorphins and tachykinins).  相似文献   

8.
The antimicrobial arenicin peptides are cationic amphipathic sequences that strongly interact with membranes. Through a cystine ring closure a cyclic β-sheet structure is formed in aqueous solution, which persists when interacting with model membranes. In order to investigate the conformation, interactions, dynamics, and topology of their bilayer-associated states, arenicin 1 and 2 were prepared by chemical solid-phase peptide synthesis or by bacterial overexpression, labeled selectively or uniformly with (15)N, reconstituted into oriented membranes, and investigated by proton-decoupled (31)P and (15)N solid-state NMR spectroscopy. Whereas the (31)P NMR spectra indicate that the peptide induces orientational disorder at the level of the phospholipid head groups, the (15)N chemical shift spectra agree well with a regular β-sheet conformation such as the one observed in micellar environments. In contrast, the data do not fit the twisted β-sheet structure found in aqueous buffer. Furthermore, the chemical shift distribution is indicative of considerable conformational and/or topological heterogeneity when at the same time the (15)N NMR spectra exclude alignments of the peptide where the β-sheet lies side ways on the membrane surface. The ensemble of experimental constraints, the amphipathic character of the peptide, and in particular the distribution of the six arginine residues are in agreement with a boatlike dimer structure, similar or related to the one observed in micellar solution, that floats on the membrane surface with the possibility to oligomerize into higher order structures and/or to insert in a transmembrane fashion.  相似文献   

9.
Upon disulfide bond reduction, the alpha 2-subunit of the dihydropyridine-sensitive Ca2+ channel undergoes a characteristic mobility shift on sodium dodecyl sulfate-polyacrylamide gel electrophoresis analysis with the concurrent appearance of the three delta peptides delta 1 (25,000 Da), delta 2 (22,000 Da), and delta 3 (17,000 Da). Densitometric scanning of Coomassie Blue-stained gels shows a stoichiometric ration of 1.0:0.31.47:0.08 for the alpha 2-subunit and the delta peptides 1, 2, and 3, respectively. Characterization of the delta peptides using antibodies, photoincorporation of a hydrophobic probe, and lectin staining shows tham to be antigenically similar hydrophobic glycoproteins. Amino-terminal sequence analysis of the delta peptides reveals three identical sequences that match the predicted amino acid sequence of the alpha 2-subunit starting at Ala935. Enzymatic deglycosylation of the reduced alpha 2.delta complex produces individual core peptides of 105,000 and 17,000 Da, respectively. Treatment of skeletal muscle membranes with high pH in the presence of reducing agents is able to extract the larger amino-terminal peptide but not the smaller carboxyl (delta) peptide, consistent with a single transmembrane domain in the carboxyl (delta) region. The data support a model of the alpha 2-subunit in which the propeptide is processed into two chains that remain attached through disulfide linkages.  相似文献   

10.
Phosphoglucose isomerase (PGI) catalyzes the isomerization of D-glucose-6-phosphate (G6P) and D-fructose-6-phosphate (F6P) in glycolysis and gluconeogenesis. Analysis of previously reported X-ray crystal structures of PGI without ligand, with the cyclic form of F6P, or with inhibitors that mimic the cis-enediol intermediate led to proposed mechanisms for the ring opening and isomerization steps in the multistep catalytic mechanism. To help complete our model of the overall mechanism, information is needed about the state of PGI between the ring opening and isomerization steps, in other words, a structure of the enzyme complexed with the open form of a substrate or an analog. Here, we report the crystal structure of rabbit PGI complexed with D-sorbitol-6-phosphate (S6P), an analog of the open chain form of G6P, at 2.0 A resolution. As was seen in the PGI/F6P structure, a helix containing amino acid residues 512-520 is found in the "out" position, which provides sufficient space in the active site for a substrate in its cyclic form and which is probably the location of that helix just after ring opening (or just before ring closure). However, the S6P ligand is in an extended conformation, as was seen previously with ligands that mimic the cis-enediol intermediate. The extended conformation enables the ligand to interact with Glu357, which transfers a proton during the isomerization step. The PGI/S6P structure represents the conformation of the enzyme and substrate between the ring opening (or ring closing) step and the isomerization step and helps to complete the model for PGI's catalytic mechanism.  相似文献   

11.
The human erythrocyte protein carboxyl methyltransferase modifies unusual protein D-aspartyl and L-isoaspartyl residues which arise spontaneously from internal rearrangements accompanying asparaginyl deamidation and aspartyl isomerization. A site of methylation associated with alpha-globin in intact cells has been identified by peptide mapping of radiolabeled globin isolated from human erythrocytes previously incubated with L-[methyl-3H]methionine. The site is located in a Staphylococcus V8 peptide containing residues 1-30 of alpha-globin. Two potential sources of methylation sites are present in this sequence at Asp-t and Asn-9.  相似文献   

12.
Asparagine (Asn) deamidation and aspartic acid (Asp) isomerization are spontaneous and common alterations occurring in pharmaceutical protein drugs in solution. Because those reactions may cause functional changes, it is important to identify the product-related substances, especially when biopharmaceuticals are under development. In this study, we used H(2)(18)O to identify Asn deamidation and Asp isomerization sites on a recombinant humanized monoclonal antibody (mAb) by using high-performance liquid chromatography-mass spectrometry (HPLC-MS). This strategy takes advantage of reactions whereby (18)O is incorporated into the protein molecule. The mAb was lyophilized and reconstituted in H(2)O or H(2)(18)O, followed by incubation at 50 degrees C for 1 month. Samples were reduced/carboxymethylated and digested by trypsin and then subjected to HPLC-MS and HPLC-tandem mass spectrometry (MS/MS) analysis. Among all of the peptide fragments analyzed, there were two in which deamidation and/or isomerization was observed. In one peptide fragment, an obvious mass shift ( approximately 3Da) at Asn was observed in the newly produced peptide when the mAb was incubated in H(2)(18)O, whereas it was barely feasible to identify this mass shift in H(2)O. In the other peptide fragment, isomerization of Asp was identified after incubation in H(2)(18)O, although it was impossible to distinguish when using H(2)O. By means of this procedure, identification of deamidation and isomerization sites can be accomplished easily even when they are difficult or impossible to detect by the usual peptide mapping.  相似文献   

13.
The cytoplasmic tail of the amyloid precursor protein (APPc) interacts with several cellular factors implicated in intracellular signaling or proteolytic production of amyloid beta peptide found in senile plaques of Alzheimer's disease patients. APPc contains two threonine residues (654 and 668 relative to APP695, or 6 and 20 relative to APPc) and a serine residue (655 or 7, respectively) that are known to be phosphorylated in vivo and may play regulatory roles in these events. We show by solution NMR spectroscopy of a 49 residue cytoplasmic tail peptide (APP-C) that in all three cases, phosphorylation induces changes in backbone dihedral angles that can be attributed to formation of local hydrogen bonds between the phosphate group and nearby amide protons. Phosphorylation of S7 also induces chemical shift changes in the hydrophobic cluster (residues I8-V13), indicating additional medium-range effects. The most pronounced changes occur upon phosphorylation of T20, a neuron-specific phosphorylation site, where the N-terminal helix capping box previously characterized for this region is altered. Characterization of torsion angles and transient hydrogen bonds indicates that prolyl isomerization of the pThr-Pro peptide bond results from both destabilization of the N-terminal helix capping box and stabilization of the cis isomer by transient hydrogen bonds. The significant population of the cis isomer (9 %) present after phosphorylation of T20 suggests a potential role of selective recognition of cis versus trans isomers in response to phosphorylation of APP. Together, these structural changes indicate that phosphorylation may act as a conformational switch in the cytoplasmic tail of APP to alter specificity and affinity of binding to cytosolic partners, particularly in response to the abnormal phosphorylation events associated with Alzheimer's disease.  相似文献   

14.
We have investigated the spontaneous degradation of aspartate and asparagine residues via succinimide intermediates in model peptides in organic co-solvents. We find that the rate of deamidation at asparagine residues is markedly reduced in solvents of low dielectric strength. Theoretical considerations suggest that this decrease in rate is due to the destabilization of the deprotonated peptide bond nitrogen anion that is the postulated attacking species in succinimide formation. This result suggests that asparagine residues in regions with low dielectric constants, such as the interior of a protein or in a membrane bilayer, are protected from this type of degradation reaction. On the other hand, we found little or no effect on the rate of succinimide-mediated isomerization of aspartate residues when subjected to the same changes in dielectric constant. In this case, the destabilization of the attacking peptide bond nitrogen anion may be balanced by increased protonation of the aspartyl side chain carboxyl group, a reaction that results in a superior leaving group. Consequently, any protein structure or conformation that would increase the protonation of an aspartate side chain carboxyl group can be expected to render that residue more labile. These results may help explain why particular aspartate residues have been found to degrade in proteins at rates comparable to those of asparagine residues, even though aspartyl-containing peptides degrade more slowly than corresponding asparaginyl-containing peptides in aqueous solutions.  相似文献   

15.
J Lowenson  S Clarke 《Blood cells》1988,14(1):103-118
As erythrocytes age in the circulation, their proteins are subjected to a wide variety of spontaneous reactions that lead to the formation of covalent derivatives. In this article, we concentrate on nonenzymatic reactions at aspartyl and asparaginyl residues, both of which are especially vulnerable targets on the protein. These residues can be altered by a combination of deamidation, isomerization, and racemization reactions that form D- and L-aspartyl and D- and L-isoaspartyl residues. We present evidence that two of these modified residues are targets for an enzymatic methyl esterification reaction, and that methylation may represent the means by which cells respond to this type of protein damage. The metabolic fate of the methyl ester is unclear, but in vitro model studies with peptides and proteins suggest that this methylation can lead to the partial repair of the altered protein and can mitigate the loss of protein function.  相似文献   

16.
The amino-acid sequences of both subunits of C-phycoerythrin from the cyanobacterium Fremyella diplosiphon have been determined. The alpha-subunit contains 164 amino acid residues, two phycoerythrobilin (PEB) chromophores and has a molecular mass of 18,368 Da (protein: 17,192 Da + 2 PEB, one PEB accounting for 588 Da). The beta-subunit consists of 184 residues, three PEB chromophores and has a molecular mass of 20,931 Da (protein: 19,168 Da and 3 PEB: 1,764 Da). The five PEB chromophores (open chain tetrapyrroles) are covalently bound to six cysteine residues (one of them doubly bound to two cysteine residues). On the alpha-subunit, the first chromophore was found at position 84, homologous to the chromophore binding site of the other biliproteins APC, PC and PEC. The second chromophore, unique for the alpha-subunit of PE, is inserted together with a pentapeptide at position 143 a. On the beta-subunit, a doubly bound chromophore is attached to cysteine residues 50 and 61, similar to the rhodophytan phycoerythrins (B-PE and R-PE). The second and third chromophores were found at positions 84 and 155, homologous to the other biliproteins. A unique peptide insertion of 14 amino acid residues (without chromophore) was found at position 141 a-o in the beta-subunit and probably is located in the three-dimensional model near the additional chromophores of the C-PE alpha- and beta-subunits. Both additional chromophores of the C-PE alpha- and beta-subunit may be located at the periphery of the C-PE-trimer. The amino-acid sequence homology between C-PE alpha- and beta-subunit is 26% and to the alpha- and beta-subunits of C-PC from Mastigocladus laminosus 49% and 48%, respectively.  相似文献   

17.
Aspartyl and asparaginyl deamidation, isomerization, and racemization reactions have been studied in synthetic peptides to model these spontaneous processes that alter protein structure and function. We show here that the peptide L-Val-L-Tyr-L-Pro-L-Asn-Gly-L-Ala undergoes a rapid deamidation reaction with a half-life of only 1.4 days at 37 degrees C, pH 7.4, to give an aspartyl succinimide product. Under these conditions, the succinimide product can further react by hydrolysis (half-time, 2.3h) and by racemization (half-time, 19.5 h). The net product of the deamidation reaction is a mixture of L- and D-normal aspartyl and beta-transpeptidation (isoaspartyl) hexapeptides. Replacement of the asparagine residue by an aspartic acid residue results in a 34-fold decrease in the rate of succinimide formation. Significant racemization was found to accompany the deamidation and isomerization reactions, and most of this could be accounted for by the rapid racemization of the succinimide intermediate. Replacement of the glycyl residue in the asparagine-containing peptide with a bulky leucyl or prolyl residue results in a 33-50-fold decrease in the rate of degradation. Peptide cleavage products are observed when these Asn-Leu and Asn-Pro-containing peptides are incubated. Our studies indicate that both aspartic acid and asparagine residues may be hot spots for the nonenzymatic degradation of proteins, especially in cells such as erythrocytes and eye lens, where these macromolecules must function for periods of about 120 days and 80 years, respectively.  相似文献   

18.
The major soluble eye lens protein, αA-crystallin, has a very long half-life. Thus, many post-translational modifications, including stereoinversion, have been found in its constituent amino acids. We determine the rates of β-linkage isomerization, which is the main reaction through the formation of a succinimide intermediate, of specific Asp residues of recombinant human αA-crystallin protein by simple RP-HPLC method. Kinetic analyses of the β-linkage isomerization were performed on the three Asp residues of αA-crystallin, (58)Asp, (84)Asp, and (151)Asp, because the d/l ratios of both the (58)Asp and (151)Asp residues were higher than 1.0 in the αA-crystallin isolated from aged human eye lens. The β-linkage isomerizations of both the (58)Asp and (84)Asp residues were suppressed in the recombinant protein by approximately 0.4-0.5 times compared to those in the synthetic peptide below 50 °C, whereas the isomerization of the (151)Asp residue occurred at the same rate for the whole protein and synthetic fragmentary peptide. The suppression of (58)Asp isomerization in the recombinant protein relaxed to some extent when the αA-crystallin protein was incubated at a high temperature. The far-UV CD spectra showed that the secondary structure of the protein was partially disordered at temperatures greater than 60 °C in the recombinant αA-crystallin protein. These results suggest that the (58)Asp residue was restrained from forming the succinimide intermediate by the higher order structure of the αA-crystallin protein, and that the structural environment around the (151)Asp residue of the αA-crystallin was similar to that of the synthetic fragmentary peptide with respect to succinimide formation. The difference in the influence of the secondary structure of the αA-crystallin protein inverts the order of the succinimide formations of the (58)Asp and (151)Asp residues in the recombinant protein as compared with the order in the synthetic fragmentary peptides.  相似文献   

19.
A NOE independent NMR method is proposed to characterize unambiguously residues involved in low populated isolated peptide helices. The method is based on the comparison of amide and H alpha chemical shift changes originated upon the addition of stabilizing or denaturing agents with true helical conformational shifts that have been measured for the first time using an isolated model peptide helix, the one formed by Ac-(Leu-Lys-Lys-Leu)3-NHEt in aqueous solution.  相似文献   

20.
The composition and distribution of rat acrosomal glycoproteins during spermiogenesis have been investigated at light and electron microscopic level by means of a variety of morphological techniques including the application of lectins conjugated to peroxidase, digoxigenin and colloidal gold, enzyme and chemical deglycosylation procedures and conventional histochemistry. Results obtained with lectin histochemistry in combination with beta-elimination reaction and endoglucosaminidase F/peptide N-glycosidase F digestion suggest that glycoproteins of mature acrosomes contain both N- and O-linked oligosaccharides. N-linked chains of acrosomal glycoproteins contain mannose and external residues of N-acetylglucosamine and galactose. They also have fucose residues linked to the core region of the oligosaccharide side chains. O-linked oligosaccharide chains contain external residues of both galactose and N-acetylgalactosamine. Mannose, fucose, galactose and N-acetylglucosamine residues were detected in acrosomes at all steps of spermiogenesis. N-acetylgalactosamine residues were only observed in the late steps of the spermiogenesis. N-acetylneuraminic acid residues were not detected throughout the acrosomal development. At initial stages of acrosome formation, glycoproteins were preferentially distributed over the acrosomic granules. In cap phase spermatids, lectin binding sites were homogeneously distributed throughout the acrosomes; however, in mature spermatozoa, glycoproteins were predominantly located over the outer acrosomal membrane.  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