首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The nonenzymatic rates of deamidation of Asn residues in a series of pentapeptides with the sequences VSNXV and VXNSV, where X is one of 10 different amino acids, were determined at neutral, alkaline, and acid pH values. The results demonstrate that in neutral and alkaline solutions the amino acid residue on the amino side of the Asn had little or no effect on the rate of deamidation regardless of its charge or size. The group on the carboxyl side of Asn affected the rate of deamidation significantly. Increasing size and branching in the side chain of this residue decreased the rate of deamidation by as much as 70-fold compared to glycine in the N-G sequence, which had the greatest rate of deamidation. In acidic solution, the rate of deamidation of the Asn residue was not affected by the amino acid sequence of the peptide. The products for each deamidation reaction were tested for the formation of isoAsp residues. In neutral and alkaline solutions, all products showed that the isoAsp:Asp peptide products were formed in about a 3:1 ratio. In acidic solution, the Asp peptide was the only deamidation product formed. All peptides in which a Ser residue follows the Asn residue were found to undergo a peptide cleavage reaction in neutral and alkaline solutions, yielding a tripeptide and a dipeptide. The rate of the cleavage reaction was about 10% of the rate of the deamidation pathway at neutral and alkaline pH values. The rates of deamidation of Asn residues in the peptides studied were not affected by ionic strength, and were not specific base catalyzed. General base catalysis was observed for small bases like ammonia. A model for the deamidation reaction is proposed to account for the observed effects.  相似文献   

2.
A study has been made on the changes in the enzymatic activity of Ribonuclease-A**-(RNase-A) exposed to highly acidic (pH less than 1) acqueous environment. Irreversible alterations of activity were observed when the protein was exposed to an acidic medium for a long period (20 to 60 h). Even prior to these changes in activity RNase-A was found to form intermediates which had very nearly the same activity as the native protein. The primary process in the acid denaturation of RNase-A was observed to be deamidation of the protein leading to the formation of active chromotographically distinct derivatives. The initial product of deamidation, a monodeamidated derivative, has been isolated by chromatography on Amberlite XE-64. This initial deamidation reaction proceeded with very high specificity. The subsequent deamidation reaction is comparatively slower, so that nearly 50% of the native protein could be converted to this derivative before any subsequent deamidation took place. This monodeamidated derivative has been designated RNase-Aa1. The conversion of RNase-A to RNase-Aa1 was not accompanied by any changes in the primary structure other than the observed deamidation. Apart from the differences in chromatographic and electrophoretic mobilities, RNase-Aa1 was found to have very nearly the same activity and physicochemical properties as the native enzyme. Significance of this specific and faster deamidation of RNase-A in this denaturing medium as well as the biological significance of such deamidation reactions of proteins are discussed.  相似文献   

3.
Transglutaminase 2 (TG2) catalyzes cross-linking or deamidation of glutamine residues in peptides and proteins. The in vivo deamidation of gliadin peptides plays an important role in the immunopathogenesis of celiac disease (CD). Although deamidation is considered to be a side-reaction occurring in the absence of suitable amines or at a low pH, a recent paper reported the selective deamidation of the small heat shock protein 20 (Hsp20), suggesting that deamidation could be a substrate dependent event. Here we have measured peptide deamidation and transamidation in the same reaction to reveal factors that affect the relative propensity for the two possible products. We report that the propensity for deamidation by TG2 is both substrate dependent and influenced by the reaction conditions. Direct deamidation is favored for poor substrates and at low concentrations of active TG2, while indirect deamidation (i.e. hydrolysis of transamidated product) can significantly contribute to the deamidation of good peptide substrates at higher enzyme concentrations. Further, we report for the first time that TG2 can hydrolyze iso-peptide bonds between two peptide substrates. This was observed also for gliadin peptides introducing a novel route for the generation of deamidated T cell epitopes in celiac disease.  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 biological function of the post-translationally methylated amino acid gamma-N-methylasparagine (gamma-NMA) in proteins is unknown. We are examining the premise that amide methylation protects against deamidation. The free amino acids Asn, gamma-NMA, Gln, and delta-N-methylglutamine (delta-NMG) were incubated at elevated temperature and a variety of pH conditions to assay for deamidation. Gln disappears 12- to 14-fold more rapidly than delta-NMG, and Asn hydrolyzes to Asp and NH3 as expected. However, the gamma-NMA deamidation rate is severely overestimated by simply measuring the disappearance of starting material because gamma-NMA undergoes a cyclization reaction in preference to deamidation. At pH 1 the predominant gamma-NMA reaction is formation of stable 3-amino-N-methylsuccinimide (NMS) and this occurs greater than 10-fold faster than Asn deamidation. At pH 4.0, 7.4, and 9.0 NMS is readily formed but it is unstable and partitions between the parent compound, gamma-NMA, and a second species, alpha-N-methylasparagine. At pH 7.4 and 9.0 gamma-NMA disappears 4-fold slower than Asn but the methyl amide hydrolysis rate is diminished by as much as 13-fold. The Asn incubations over the pH range 1-9 yield scant evidence of a succinimide intermediate. It is concluded that the amide methylation provides a unique reaction pathway and stabilization for the N-methylsuccinimide species. Amino acid analysis by o-phthalaldehyde postcolumn reaction fails to detect isoasparagine, alpha-N-methylasparagine, and NMS. Amino acid analysis by precolumn derivatization with phenyl isothiocyanate destroys NMS and therefore cannot quantitate this compound. The ninhydrin postcolumn derivatization method is able to detect and quantitate all of these amino acid species.  相似文献   

6.
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   

7.
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   

8.
Selective deamidation of Asn67 of RNase A to beta-Asp67 and Asp67 residues at neutral pH initially produces greater amounts of the beta-Asp derivative. As the reaction proceeds the relative concentration of [Asp67]-RNase A increases and, at equilibrium, becomes predominant. Such a discrepancy between the kinetic and thermodynamic control on reaction products is discussed in light of information from X-ray three-dimensional analysis and the lower thermodynamic stability of the beta-Asp derivative relative to the parent enzyme.  相似文献   

9.
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   

10.
The primary sequence dependence of deamidation has been quantitatively explained on the basis of a simple steric and catalytic model. Application to the known deamidation rates of peptides produces a table of coefficients that permits calculation of the known deamidation rates and prediction of deamidation rates for peptide sequences that have not yet been measured. This work permits a better understanding of deamidation, provides a prediction procedure for protein engineering, and facilitates improved computation of peptide and protein primary, secondary, tertiary, and quaternary structure deamidation rates.  相似文献   

11.
Tissue transglutaminase (TG2) can modify proteins by transamidation or deamidation of specific glutamine residues. TG2 has a major role in the pathogenesis of celiac disease as it is both the target of disease-specific autoantibodies and generates deamidated gliadin peptides that are recognized by CD4(+), DQ2-restricted T cells from the celiac lesions. Capillary electrophoresis with fluorescence-labeled gliadin peptides was used to separate and quantify deamidated and transamidated products. In a competition assay, the affinity of TG2 to a set of overlapping gamma-gliadin peptides was measured and compared with their recognition by celiac lesion T cells. Peptides differed considerably in their competition efficiency. Those peptides recognized by intestinal T cell lines showed marked competition indicating them as excellent substrates for TG2. The enzyme fine specificity of TG2 was characterized by synthetic peptide libraries and mass spectrometry. Residues in positions -1, +1, +2, and +3 relative to the targeted glutamine residue influenced the enzyme activity, and proline in position +2 had a particularly positive effect. The characterized sequence specificity of TG2 explained the variation between peptides as TG2 substrates indicating that the enzyme is involved in the selection of gluten T cell epitopes. The enzyme is mainly localized extracellularly in the small intestine where primary amines as substrates for the competing transamidation reaction are present. The deamidation could possibly take place in this compartment as an excess of primary amines did not completely inhibit deamidation of gluten peptides at pH 7.3. However, lowering of the pH decreased the reaction rate of the TG2-catalyzed transamidation, whereas the rate of the deamidation reaction was considerably increased. This suggests that the deamidation of gluten peptides by TG2 more likely takes place in slightly acidic environments.  相似文献   

12.
The phosphorylation of the primary gene products of alpha-crystallin   总被引:1,自引:0,他引:1  
The alpha-crystallin primary gene product A2 and its post-translational modified counterpart A1 were isolated from calf lens cortex. The amino acid compositions determined from both chains were almost identical and in excellent agreement with that calculated from the reported sequence of A2. Chemical analysis of phosphate revealed 1 mol/mol of A1 and was negative in A2. Phosphoamino acid analysis demonstrated the presence of phosphoserine only in A1. Chymotryptic peptide maps of A2 and A1 resolved approximately 50 peptides and were strikingly similar. An apparent change in the relative mobility of one peptide was the only difference observed between A1 and A2. Phosphate analysis of this peptide obtained from A1 and A2 was positive only in the peptide from A1. Identical amino acid composition and the sequence Arg-Leu-Pro-Ser-Asn-Val-Asp-Gln-Ser-Ala-Leu was found for the peptide isolated from both chains, corresponding to residues 119 to 129 in the reported sequence of A2. These results indicate that the post-translational modification of A2 to A1 is the result of a phosphorylation reaction rather than a spontaneous nonenzymatic deamidation as previously suggested.  相似文献   

13.
Two dimensional electrophoresis has revealed a microheterogeneity in the recombinant human phenylalanine hydroxylase (hPAH) protomer, that is the result of spontaneous nonenzymatic deamidations of labile asparagine (Asn) residues [Solstad, T. and Flatmark, T. (2000) Eur. J. Biochem.267, 6302-6310]. Using of a computer algorithm, the relative deamidation rates of all Asn residues in hPAH have been predicted, and we here verify that Asn32, followed by a glycine residue, as well as Asn28 and Asn30 in a loop region of the N-terminal autoregulatory sequence (residues 19-33) of wt-hPAH, are among the susceptible residues. First, on MALDI-TOF mass spectrometry of the 24 h expressed enzyme, the E. coli 28-residue peptide, L15-K42 (containing three Asn residues), was recovered with four monoisotopic mass numbers (i.e., m/z of 3106.455, 3107.470, 3108.474 and 3109.476, of decreasing intensity) that differed by 1 Da. Secondly, by reverse-phase chromatography, isoaspartyl (isoAsp) was demonstrated in this 28-residue peptide by its methylation by protein-l-isoaspartic acid O-methyltransferase (PIMT; EC 2.1.1.77). Thirdly, on incubation at pH 7.0 and 37 degrees C of the phosphorylated form (at Ser16) of this 28-residue peptide, a time-dependent mobility shift from tR approximately 34 min to approximately 31 min (i.e., to a more hydrophilic position) was observed on reverse-phase chromatography, and the recovery of the tR approximately 34 min species decreased with a biphasic time-course with t0.5-values of 1.9 and 6.2 days. The fastest rate is compatible with the rate determined for the sequence-controlled deamidation of Asn32 (in a pentapeptide without 3D structural interference), i.e., a deamidation half-time of approximately 1.5 days in 150 mm Tris/HCl, pH 7.0 at 37 degrees C. Asn32 is located in a cluster of three Asn residues (Asn28, Asn30 and Asn32) of a loop structure stabilized by a hydrogen-bond network. Deamidation of Asn32 introduces a negative charge and a partial beta-isomerization (isoAsp), which is predicted to result in a change in the backbone conformation of the loop structure and a repositioning of the autoregulatory sequence and thus affect its regulatory properties. The functional implications of this deamidation was further studied by site-directed mutagenesis, and the mutant form (Asn32-->Asp) revealed a 1.7-fold increase in the catalytic efficiency, an increased affinity and positive cooperativity of L-Phe binding as well as substrate inhibition.  相似文献   

14.
The effects of secondary structure on asparagine (N) deamidation in a 22 amino acid sequence (369‐GFYPSDIAVEWESNGQPENNYK‐390) of the crystallizable (Fc) fragment of a human monoclonal antibody (Fc IgG1) were investigated using high‐resolution ultra performance liquid chromatography with tandem mass spectrometry (UPLC/MS). Samples containing either the intact Fc IgG (~50 kD) (“intact protein”), or corresponding synthetic peptides (“peptide”) were stored in Tris buffer at 37°C and pH 7.5 for up to forty days, then subjected to UPLC/MS analysis with high energy MS1 fragmentation. The peptide deamidated only at N382 to form the isoaspartate (isoD382) and aspartate (D382) products in the ratio of ~4:1, with a half‐life of ~3.4 days. The succinimide intermediate (Su382) was also detected; deamidation was not observed for the other two sites (N387 and N388) in peptide samples. The intact protein showed a 30‐fold slower overall deamidation half‐life of ~108 days to produce the isoD382 and D387 products, together with minor amounts of D382. Surprisingly, the D382 and isoD387 products were not detected in intact protein samples and, as in the peptide samples, deamidation was not detected at N388. The results indicate that higher order structure influences both the rate of N‐deamidation and the product distribution.  相似文献   

15.
Why does ribonuclease irreversibly inactivate at high temperatures?   总被引:13,自引:0,他引:13  
S E Zale  A M Klibanov 《Biochemistry》1986,25(19):5432-5444
The mechanism of irreversible thermoinactivation of bovine pancreatic ribonuclease A in the pH range relevant to enzymatic catalysis has been elucidated. At 90 degrees C and pH 4, the enzyme inactivation is caused by hydrolysis of peptide bonds at aspartic acid residues (the main process) and deamidation of asparagine and/or glutamine residues. At 90 degrees C and neutral pH (pH 6 and 8), the enzyme inactivation is caused by a combination of disulfide interchange (the main process), beta-elimination of cystine residues, and deamidation of asparagine and/or glutamine residues. These four processes appear to demarcate the upper limit of thermostability of enzymes.  相似文献   

16.
Anticariogenic casein phosphopeptides (ACPP) contain the cluster sequence -Ser(P)-Ser(P)-Ser(P)-Glu-Glu- and have commercial potential as toothpaste, mouthwash, and food additives for the prevention of dental caries. In an approach to develop a commercial-scale process for the production of ACPP we have comprehensively characterized casein phosphopeptides (CPP) produced under industrially relevant conditions. Sodium caseinate (10% w/v) was hydrolyzed by Novo trypsin (commercial grade) at 50 degrees C for 2 h and CPP were purified from the acid clarified hydrolysate by a single-step selective precipitation procedure involving Ca(2+) (20 mol/mol casein) and ethanol (50% v/v) at pH 4.6 or 8.0. The individual peptides of the CPP preparations were purified by reversed-phase high-performance liquid chromatography (HPLC) and then identified by amino acid composition and sequence analyses. The yield of the pH 8.0 precipitate (13.85 +/- 0.48 wt % of the caseinate) was slightly higher than that of the pH 4.6 precipitate (11.04 +/- 0.30 wt % of the caseinate). However, the pH 4.6 precipitate contained predominantly (86.4 mol %) ACPP cluster peptides with small amounts of the diphosphorylated peptides (13.6 mol %), alpha(s1)(43-58) and alpha(s2)(126-136). In the pH 8.0 precipitate the cluster peptides represented a smaller proportion of the total peptides (61.9 mol %) due to increased recoveries of the diphosphorylated peptides (24.4 mol %) as well as the additional recovery of the monophosphorylated peptide beta(33-48) (13.7 mol %) indicating increased cross-linking by Ca(2+) at the higher pH. The recovery of the ACPP from the original caseinate was similar for both the pH 4.6 and 8.0 precipitates. Slight chymotryptic activity was detected in the industrial-grade enzyme, resulting in minor truncation of some peptides. Also some deamidation and methionine oxidation of one peptide, alpha(s1)(59-79), were detected. In conclusion, ACPP can be produced under industrially relevant conditions with only minor modifications such as slight truncation, deamidation, and methionine oxidation. However, in order to prepare casein phosphopeptides predominantly containing the cluster sequence -Ser(P)-Ser(P)-Ser(P)-Glu-Glu-, the single-step selective precipitation with Ca(2+)/ethanol should be performed at pH 4.6 rather than pH 8.0. (c) 1995 John Wiley & Sons, Inc.  相似文献   

17.
Identification of deamidated sites in proteins is commonly used for assignment of N-glycosylation sites. It is also important for assessing the role of deamidation in vivo. However, nonenzymatic deamidation occurs easily in peptides under conditions commonly used in treatment with trypsin and PNGase F. The impact on proteomic sample preparation has not yet been evaluated systematically. In addition, the (13)C peaks of amidated peptides can be misassigned as monoisotopic peaks of the corresponding deamidated ones in database searches. The 19.34 mDa mass difference between them is proposed as a means for eliminating the resulting false positive identifications in large-scale proteomic analysis. We evaluated five groups of proteomic data, obtained mainly through an electrostatic repulsion-hydrophilic interaction chromatography (ERLIC)-reverse phase (RP) chromatography sequence, and ascertained that nonenzymatic asparagine deamidation occurred to some extent on 4-9% of the peptides, resulting in the false positive identification of many N-glycosylation sites. A comprehensive investigation indicated that the chief causative factors were the mildly alkaline pH and prolonged incubations at 37 °C during proteomic sample preparation. An improved protocol is proposed featuring tryptic digestion at pH 6 and deglycosylation at pH 5, resulting in a significant decrease in nonenzymatic deamidation while conserving adequate digestion efficiency. The number of identified deamidation sites was improved significantly by increasing the sample loading amount in liquid chromatography-tandem MS. This permitted the identification of a significant number of glutamine deamidation sites, which featured sequence motifs largely different from those for asparagine deamidation: -Q-V-, -Q-L- and -Q-G- and, to a lesser extent, -Q-A- and -Q-E-.  相似文献   

18.
Asn deamidation was monitored in Ala-based octadecapeptides of varying alpha-helicity. Gly was substituted for Ala residues at positions 6 and 16 to create a peptide with less helicity. Ala --> Gly substitutions were made at three or more residues from the Asn to negate known primary sequence effects on deamidation rates. The extent of helicity and rate of Asn deamidation for alkaline aqueous solutions of each peptide was measured as a function of temperature by circular dichroism and reversed-phase high-performance liquid chromatography, respectively. The rate of deamidation in the peptides was inversely proportional to the extent of alpha-helicity. The results support the conclusion that Asn deamidation only occurs in the nonhelical population of conformers.  相似文献   

19.
The 'protein world' exhibits additional complexity caused by post-translational modifications. One such process is nonenzymic deamidation of asparagine which is controlled partly by primary sequence, but also higher order protein structure. We have studied the deamidation of an N-terminal peptide in muscle glyceraldehyde 3-phosphate dehydrogenase to relate three-dimensional structure, proteolysis, and deamidation. This work has significant consequences for identification of proteins using peptide mass fingerprinting.  相似文献   

20.
The rates of deamidation of Asn and Gln residues in peptides and proteins depend upon both the identity of other nearby amino acid residues, some of which can catalyze the deamidation reaction of the Asn and Gln side chains, and upon polypeptide conformation. Proximal amino acids can be contiguous in sequence or brought close to Asn or Gln side chains by higher order structure of the protein. Local polypeptide conformation can stabilize the oxyanion transition state of the deamidation reaction and also enable deamidation through the beta-aspartyl shift mechanism. In this paper, the environments of Asn and Gln residues in known protein structures are examined to determine the configuration and identity of groups which participate in deamidation reactions. Sequence information is also analyzed and shown to support evolutionary selection against the occurrence of certain potentially catalytic amino acids adjacent to Asn and Gln in proteins. This negative selection supports a functional role for deamidation in those non-mutant proteins in which it occurs.  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