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.
Solid-phase peptide synthesis and deamidation measurements using a novel mass spectrometric technique were carried out for 94 model asparaginyl peptides from 3 to 13 residues in length. Deamidation rates of these peptides in pH 7.4, 37.0 degrees C, 0.15 M Tris-HCl buffer were measured and evaluated. It was found that they validate the use of pentapeptide models as surrogates for the primary sequence dependence of peptide and protein deamidation rates and the discovery by difference of secondary, tertiary and quaternary structure effects. Deamidation of the pentapeptide models, compared with that of longer peptides of more intricate structure, is discussed, and the application of this technique to deamidation measurement of intact proteins is demonstrated.  相似文献   

2.
Deamidation of asparaginyl residues in proteins produces a mixture of asparaginyl, n-aspartyl, and isoaspartyl residues, which affects the proteins' structure, function, and stability. Thus, it is important to identify and quantify the products to evaluate the effects in biological systems. It is still a challenging task to distinguish between the n-Asp and isoAsp deamidation products in a proteome-wide analysis because of their similar physicochemical properties. The quantification of the isomeric deamidated peptides is also rather difficult because of their coelution/poor separation in reverse-phase liquid chromatography (RPLC). We here propose a RP-ERLIC-MS/MS approach for separating and quantifying on a proteome-wide scale the three products related to deamidation of the same peptide. The key to the method is the use of RPLC in the first dimensional separation and ERLIC (electrostatic repulsion-hydrophilic interaction chromatography) in the second, with direct online coupling to tandem MS. The coelution of the three deamidation-related peptides in RPLC is then an asset, as they are collected in the same fraction. They are then separated and identified in the second dimension with ERLIC, which separates peptides on the basis of both pI and GRAVY values. The coelution of the three products in RPLC and their efficient separation in ERLIC were validated using synthetic peptides, and the performance of ERLIC-MS/MS was tested using peptide mixtures from two proteins. Applying this sequence to rat liver tissue, we identified 302 unique N-deamidated peptides, of which 20 were identified via all three deamidation-related products and 70 of which were identified via two of them.  相似文献   

3.
Nonenzymatic intramolecular reactions can result in the deamidation, isomerization, and racemization of protein and peptide asparaginyl and aspartyl residues via succinimide intermediates. To understand the sequence dependence of these reactions, we measured the rate of succinimide formation in a series of synthetic peptides at pH 7.4. These peptides (Val-Tyr-Pro-X-Y-Ala) contained an internal aspartyl, asparaginyl, aspartyl beta-methyl ester, or aspartyl alpha-methyl ester residue (X) followed by a glycyl, seryl, or alanyl residue (Y). The rates of succinimide formation of the asparaginyl peptides were found to be 13.1-35.6 times faster than those of the aspartyl peptides. The rates of succinimide formation for the glycyl peptides were 6.5-17.6 times faster than those of the alanyl peptides, while the rates for the seryl peptides were 1.6-4.5 times faster than those of the alanyl peptides. The overall 232-fold range in these reaction rates for aspartyl and asparaginyl residues suggests that sequence can be an important determinant in their stability in flexible peptides. In proteins, there may be a much larger range in the rates of succinimide formation because specific conformations may greatly enhance or inhibit this reaction.  相似文献   

4.
R Lura  V Schirch 《Biochemistry》1988,27(20):7671-7677
The tetrapeptides Val-Asn-Gly-Ala and N-acetyl-Val-Asn-Gly-Ala undergo deamidation of the asparaginyl residue at pH 7.0 at similar rates. However, they form different products. The N-acetyl peptide gave a 3:1 ratio of N-acetyl-Val-isoAsp-Gly-Ala and N-acetyl-Val-Asp-Gly-Ala, respectively. The nonacetylated peptide gave no detectable amounts of these products but rather gave a cyclic peptide formed from the nucleophilic displacement of the asparaginyl side chain amide by the amino terminus of valine. This compound was slowly inverted at carbon 2 of the asparaginyl residue. At pH values above 7.5, the nonacetylated peptide also underwent deamidation to form Val-isoAsp-Gly-Ala and Val-Asp-Gly-Ala in the 3:1 ratio. Proton NMR spectra of the acetylated and nonacetylated tetrapeptides show that below pH 7.5 they have very different preferred conformations, and it is these different conformations which result in the different mechanisms of deamidation. Above pH 9.0, both peptides have similar conformations and deamidate by the same mechanism to give equivalent products. Neither mechanism of deamidation was subject to general base catalysis by the buffer. These results suggest that deamidation rates of the asparaginyl-glycyl sequence in proteins will vary according to the conformation of the peptide backbone of each respective protein. The results also show that asparaginyl residues which are penultimate to the amino terminus can react to form an N-terminal-blocked seven-membered ring.  相似文献   

5.
The non-enzymatic deamidation of asparaginyl residues is a major source of spontaneous damage of several proteins under physiological conditions. In many cases, deamidation and isoaspartyl formation alters the biological activity or stability of the native polypeptide. Rates of deamidation of particular residues depend on many factors including protein structure and solvent exposure. Here, we investigated the spontaneous deamidation of the two NADP-glutamate dehydrogenase isoenzymes from Saccharomyces cerevisiae, which have different kinetic properties and are differentially expressed in this yeast. Our results show that Asn54, present in Gdh3p but missing in the GDH1-encoded homologue, is readily deamidated in vitro under alkaline conditions. Relative to the native enzyme, deamidated Gdh3p shows reduced protein stability. The different deamidation rates of the two isoenzymes could explain to some extent, the relative in vivo instability of the allosteric Gdh3p enzyme, compared to that of Gdh1p. It is thus possible that spontaneous asparaginyl modification could play a role in the metabolic regulation of ammonium assimilation and glutamate biosynthesis.  相似文献   

6.
Rabbit liver cytosolic serine hydroxymethyltransferase exists in several subforms which have different isoelectric points. Incubation of the purified enzyme with chymotrypsin cleaves the enzyme at Trp14. The released amino-terminal 14-mer peptide was shown to exist in three forms of equal concentration. The peptides differ in structure only at the asparaginyl residue at position 5. In addition to asparagine at this position we found both aspartyl and isoaspartyl residues. The deamidation of Asn5 does not appear to occur during the purification of the enzyme. The in vitro rate of deamidation of Asn5 in the enzyme is more than 5-fold slower than the rate of deamidation of this residue in the free 14-mer peptide. The isoaspartyl residue at position 5 serves as a substrate for protein carboxyl methyltransferase both in the free 14-mer peptide and the native enzyme. The enzyme which has had the amino-terminal 14 residues removed by digestion with chymotrypsin still exists in several forms with different isoelectric points. Reaction of peptides from this enzyme with carboxyl methyltransferase suggests that there is at least one more asparaginyl residue in this enzyme other than Asn5 which has undergone deamidation with the formation of isoaspartyl bonds.  相似文献   

7.
Spontaneous isoaspartyl formation from aspartyl dehydration or asparaginyl deamidation is a major source of modifications in protein structures. In cells, these conformational changes could be reverted by the protein L-isoaspartyl methyltransferase (PIMT) repair enzyme that converts the isoaspartyl residues into aspartyl. The physiological importance of this metabolism has been recently illustrated in plants. Recent developments allowing peptide isomer identification and quantification at the proteome scale are portrayed. The relevance of these new proteomic approaches based on 2-D electrophoresis or electron capture dissociation analysis methods was initially documented in mammals. Extended use to Arabidopsis model systems is promising for the discovery of controlling mechanisms induced by these particular post-translational modifications and their biological role in plants.  相似文献   

8.
We have shown the dependence of the deamidation half-times of the peptides, GlyLeuGlnAlaGly and GlyArgGlnAlaGly upon pH, temperature, and ionic strength. Increase in temperature or ionic strength, variation of pH to pH′s higher or lower than pH 6, and the use of phosphate buffer rather than Tris buffer at high pH all decrease the half-time of dcamidation. Temperature increase of 20°C or pH change of 2 pH units decreases the half-time about fivefold, while increase of one ionic strength unit decreases the half-time about twofold. In pH 7.4, I = 0.2, 37.0°C phosphate buffer, the deamidation half-times are 663 ± 74 and 389 ± 56 days respectively for the two peptides, GlyLeuGlnAlaGly and GlyArgGlnAlaGly.These experiments should serve as a warning to peptide and protein experimenters that even the more stable glutaminyl residues are unstable with respect to deamidation in certain solvent conditions. These experiments also provide, along with previously reported experiments on asparaginyl peptides (7), some quantitative data to help with the extrapolation of in vitro deamidation experiments to in vivo deamidation conditions.  相似文献   

9.
Nonenzymatic deamidation rates for 52 glutaminyl and 52 asparaginyl pentapeptides in pH 7.4, 37.0 degrees C. 0.15 m Tris-HCl buffer have been determined by direct injection mass spectrometry. These and the previously reported 306 asparginyl rates have been combined in a self-consistent model for peptide deamidation. This model depends quantitatively upon peptide structure and involves succinimide, glutarimide and hydrolysis mechanisms. The experimental values and suitable interpolated values have been combined to provide deamidation rate values in pH 7.4, 37.0 degrees C. 0.15 m Tris-HCl buffer for the entire set of 648 single-amide permutations of ordinary amino acid residues in GlyXxxAsnYyyGly and GlyXxxGlnYyyGly. Thus, knowledge about sequence-dependent deamidation in peptides is extended to include very long deamidation half-times in the range of 2-50 years.  相似文献   

10.
11.
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   

12.
Deamidation of asparaginyl residues is a common posttranslational modification in proteins and has been studied extensively because of its important biological effects, such as those on enzymatic activity, protein folding, and proteolytic degradation. However, characterization of the sites of deamidation of a protein has been a difficult analytical problem. In this study, mass spectrometry has been used as an analytical tool to characterize the deamidation of barstar, an RNAse inhibitor. Upon incubation of the protein at alkaline pH for 5 h, intact mass analysis of barstar, using electrospray ionization quadrupole time-of-flight mass spectrometry (ESI QToF MS), indicated an increase in the mass of +2 Da, suggesting possible deamidation of the protein. The sites of deamidation have been identified using the conventional bottom-up approach using a capillary liquid chromatography connected on line to an ESI QToF mass spectrometer and top down approach by direct infusion of the intact protein and fragmenting inside MS. These chemical modifications are shown to lead to stabilization of an unfolding intermediate, which can be observed in equilibrium unfolding studies.  相似文献   

13.
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   

14.
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   

15.
The influence of some hydrocarbons that are often used at different stages of immunobiological preparation's production as stabilizers of biological activity on the dynamics of nonenzymatic deamidation in proteins of immune whey against conditionally pathogenic microorganisms obtained by means of membrane ultrafiltration technology is investigated. Preparations of whey were incubated in 10 per cent solutions of glucose, fructose and sorbitol at the conditions similar to physiological ones (0.9% NaCl, pH 5.5) and temperature of about +4 degrees C and +35 degrees C for 7, 14 and 28 days. A sample dissolved in 0.9% NaCl (pH 5.5) without addition of hydrocarbons was used as a "control preparate". All explored substances brought about the suppressive effect on deamidation rate of asparaginyl residues whereas that of glutaminyl residues, on the contrary, was obviously increased. The possible reasons for these observations are discussed.  相似文献   

16.
The cellular concentration of Bcl-xL is among the most important determinants of treatment response and overall prognosis in a broad range of tumors as well as an important determinant of the cellular response to several forms of tissue injury. We and others have previously shown that human Bcl-xL undergoes deamidation at two asparaginyl residues and that DNA-damaging antineoplastic agents as well as other stimuli can increase the rate of deamidation. Deamidation results in the replacement of asparginyl residues with aspartyl or isoaspartyl residues. Thus deamidation, like phosphorylation, introduces a negative charge into proteins. Here we show that the level of human Bcl-xL is constantly modulated by deamidation because deamidation, like phosphorylation in other proteins, activates a conditional PEST sequence to target Bcl-xL for degradation. Additionally, we show that degradation of deamidated Bcl-xL is mediated at least in part by calpain. Notably, we present sequence and biochemical data that suggest that deamidation has been conserved from the simplest extant metazoans through the human form of Bcl-xL, underscoring its importance in Bcl-xL regulation. Our findings strongly suggest that deamidation-regulated Bcl-xL degradation is an important component of the cellular rheostat that determines susceptibility to DNA-damaging agents and other death stimuli.  相似文献   

17.
Mimetics of beta-turn structures in proteins have been used to calibrate the relative reactivities toward deamidation of asparagine residues in the two central positions of a beta-turn and in a random coil. N-Acetyl-Asn-Gly-6-aminocaproic acid, an acyclic analog of a beta-turn mimic undergoes deamidation of the asparaginyl residue through a succinimide intermediate to generate N-acetyl-Asp-N-Gly-6-aminocaproic acid (6-aminocaproic acid, hereafter Aca) and N-acetyl-L-iso-aspartyl (isoAsp)-Gly-Aca (pH 8.8, 37 degrees C) approximately 3-fold faster than does the cyclic beta-turn mimic cyclo-[L-Asn-Gly-Aca] with asparagine at position 2 of the beta-turn. The latter compound, in turn, undergoes deamidation approximately 30-fold faster than its positional isomer cyclo-[Gly-Asn-Aca] with asparagine at position 3 of the beta-turn. Both cyclic peptides assume predominantly beta-turn structures in solution, as demonstrated by NMR and circular dichroism characterization. The open-chain compound and its isomer N-acetyl-Gly-Asn-Aca assume predominantly random coil structures. The latter isomer undergoes deamidation 2-fold slower than the former. Thus the order of reactivity toward deamidation is: asparagine in a random coil approximately 3x(asparagine) in position 2 of a beta-turn approximately 30x (asparagine) in position 3 of a beta-turn.  相似文献   

18.
Deamidation, the loss of the ammonium group of asparagine and glutamine to form aspartic and glutamic acid, is one of the most commonly occurring post-translational modifications in proteins. Since deamidation rates are encoded in the protein structure, it has been proposed that they can serve as molecular clocks for the timing of biological processes such as protein turnover, development and aging. Despite the importance of this process, there is a lack of detailed structural information explaining the effects of deamidation on the structure of proteins. Here, we studied the effects of deamidation on human triosephosphate isomerase (HsTIM), an enzyme for which deamidation of N15 and N71 has been long recognized as the signal for terminal marking of the protein. Deamidation was mimicked by site directed mutagenesis; thus, three mutants of HsTIM (N15D, N71D and N15D/N71D) were characterized. The results show that the N71D mutant resembles, structurally and functionally, the wild type enzyme. In contrast, the N15D mutant displays all the detrimental effects related to deamidation. The N15D/N71D mutant shows only minor additional effects when compared with the N15D mutation, supporting that deamidation of N71 induces negligible effects. The crystal structures show that, in contrast to the N71D mutant, where minimal alterations are observed, the N15D mutation forms new interactions that perturb the structure of loop 1 and loop 3, both critical components of the catalytic site and the interface of HsTIM. Based on a phylogenetic analysis of TIM sequences, we propose the conservation of this mechanism for mammalian TIMs.  相似文献   

19.
Smearing from high-molecular-mass regions to low-molecular-mass regions on western blot is the most striking observation of the tau making up paired helical filaments in brain tissues affected by Alzheimer's disease. Because our previous study showed site-specific deamidation/isomerization in the smeared tau in vivo, a feature of protein aging, recombinant tau was subjected to prolonged (up to 90 days) in vitro incubation. Carboxymethylated tau at approximately 50 kDa gradually disappeared and was converted to dimers and to high- and low-molecular-mass smearing. In addition, the same site-specific deamidation/isomerization as previously identified in the smeared tau in vivo emerged. Most importantly, tau was spontaneously degraded, generating fragments that start from bulky residues next to asparaginyl residues. This spontaneous degradation of tau probably represents non-enzymatic cleavage through the formation of succinimide intermediates. Similar degradation products starting from the bulky residues next to asparaginyl residues were found in the smeared tau in vivo partially purified from the homogenates from Alzheimer's disease brains.  相似文献   

20.
Three chemical specific cleavage reactions, one for the carboxyl side of aspartyl peptide bonds, one for the carboxyl side of asparaginyl peptide bonds and another for the amino side of seryl/threonyl peptide bonds have been recently established. Additionally, these reactions simultaneously react on several post-translationally modified groups in peptides or proteins. The modified groups cover the external modifications N-formyl, N-acetyl, N-pyroglutamyi residues and C-terminal-alpha amide, as well as the internal modifications such as O-acetyl serine, phosphorylated serine/tyrosine, sulfonylated tyrosine, glycosylated serine/threonine and glycosylated asparagine. These three cleavage reactions relate to key amino acids for modifications, deamidation for asparagine, phosphorylation and acetylation for serine, and glycosylation for asparagine, serine and threonine. The chemical reactions on these modifications change the peptide mapping pattern, and information from these reactions may contribute characterization and location of post-translational modified groups in the protein.  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