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 31 毫秒
1.
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   

2.
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   

3.
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   

4.
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   

5.
6.
The effect of lysine residues on the deamidation reaction of the asparagine side chain has been studied on the peptide and on its lysine-acetylated derivative in a wide range of pH values. The amino acid sequence of these peptides is similar to the local sequence flanking the labile Asn-67 in RNAse A. The experimental data show that Lys influences both the deamidation rate and the relative yield of the two reaction products, i.e., the aspartic acid and beta-aspartic acid containing peptide. These effects are pH dependent and can be rationalized based on the mechanism previously proposed for the deamidation reaction via succinimide derivative.  相似文献   

7.
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   

8.
The degradation of proteins by asparagine deamidation and aspartate isomerization is one of several chemical degradation pathways for recombinant antibodies. In this study, we have identified two solvent accessible degradation sites (light chain aspartate-56 and heavy chain aspartate-99/101) in the complementary-determining regions of a recombinant IgG1 antibody susceptible to isomerization under elevated temperature conditions. For both hot-spots, the degree of isomerization was found to be significantly higher than the deamidation of asparagine-(387, 392, 393) in the conserved CH3 region, which has been identified as being solvent accessible and sensitive to chemical degradation in previous studies. In order to reduce the time for simultaneous identification and functional evaluation of potential asparagine deamidation and aspartate isomerization sites, a test system employing accelerated temperature conditions and proteolytic peptide mapping combined with quantitative UPLC-MS was developed. This method occupies the formulation buffer system histidine/HCl (20 mM; pH 6.0) for denaturation/reduction/digestion and eliminates the alkylation step. The achieved degree of asparagine deamidation and aspartate isomerization was adequate to identify the functional consequence by binding studies. In summary, the here presented approach greatly facilitates the evaluation of fermentation, purification, formulation, and storage conditions on antibody asparagine deamidation and aspartate isomerization by monitoring susceptible marker peptides located in the complementary-determining regions of recombinant antibodies.  相似文献   

9.
Mammalian protein carboxyl methyltransferases have recently been proposed to recognize atypical configurations of aspartic acid and may possibly function in the metabolism of covalently altered cellular proteins. Consistent with this proposal, the tetrapeptide tetragastrin, containing a single "normal" L-aspartyl residue (L-Trp-L-Met-L-Asp-L-Phe-NH2) was found here not to be an in vitro substrate for erythrocyte carboxyl methyltransferase activity. However, chemical treatment of tetragastrin by methyl esterification and then de-esterification of the aspartic acid residue yielded a mixture of peptide products, the major one of which could now be enzymatically methylated. We show here that this new peptide species is the isomeric beta-aspartyl form of tetragastrin (L-iso-tetragastrin; L-Trp-L-Met-L-Asp-L-Phe-NH2), and it appears that isomerization proceeds via an intramolecular succinimide intermediate during the de-esterification procedure. L-iso-Tetragastrin is stoichiometrically methylated (up to 90% in these experiments) with a Km for the enzyme of 5.0 microM. Similar chemical treatment of several other L-aspartyl peptides also resulted in the formation of new methyltransferase substrates. This general method for converting normal aspartyl peptides to isoaspartyl peptides may have application in the reverse process as well.  相似文献   

10.
Site-specific racemization in aging alpha A-crystallin   总被引:1,自引:0,他引:1  
Of all aspartyl residues in bovine alpha A-crystallin, only Asp-151 exhibits pronounced racemization. Asp-151 is also one of the sites where peptide bond cleavage occurs in in vivo aging alpha A-crystallin. This aspartyl residue is followed by an alanyl residue and resides in a flexible carboxyl terminal extension of alpha-crystallin. Both in vivo and in vitro racemization studies indicate that the pronounced and site-specific racemization of Asp-151 proceeds via formation of a succinimide intermediate. The in vivo racemization of aspartyl residues in alpha A-crystallin is discussed with regard to the proposed tertiary structure of alpha-crystallin.  相似文献   

11.
The age-related formation of succinimides in proteins, through spontaneous deamidation of asparagine, and through cyclization of aspartic acid, is thought to be followed by the hydrolysis of the succinimide ring, yielding a mixture of normal aspartic acid sites and-isomerized aspartic acid sites (isoaspartic acid). The chemical reduction of an isoaspartyl site to the corresponding amino acid alcohol, isohomoserine, has now been investigated as a general approach to measuring the accumulation of isomerized residues in aging proteins. The methods employed were based on conditions previously found to be successful in reducing protein aspartic acid to homoserine. Borane was employed as the reducing agent, and was found to produce the expected amino acid alcohols in reactions with model peptides. In addition, amino acid analysis revealed a complex pattern of unknown products of these reduction reactions, some of which were also evident when a much stronger reducing agent, lithium aluminum hydride, was used. The correlation of some of these side-products with the isomerization of the peptide suggests, unexpectedly, that the reactivity of reducing agents toward aspartyl residues and perhaps other sites in the peptide may be influenced by steric factors related to aspartyl isomerization. The borane reduction method was also applied to proteins. No detectable isohomoserine was formed either in ovalbumin, a model aged protein, or in human lens proteins of advanced age, with conditions that fully reduced normal aspartyl residues to homoserine. These tests thus indicate that the percentage of aspartic acid in the isomerized form in these proteins is below the limit of detectability (below 5%). These results complement previous experimental results that have indicated a low bulk isoaspartyl content in most natural proteins.  相似文献   

12.
The substrate specificity of an intracellular proteinase from Streptococcus lactis was investigated in an effort to understand the role of the enzyme in the cell. Peptides in which the N-terminal residue was glycine were not hydrolyzed by the enzyme (exceptions were glycyl-alanine, glycyl-aspartic acid, and glycyl-asparagine), but the peptide was hydrolyzed if the N-terminal residue was alanine. The enzyme also showed activity toward peptides containing aspartic acid or asparagine. Hydrolysis of only the peptide bonds of alanyl, aspartyl, or asparaginyl residues was confirmed by the action of the enzyme on oxidized bovine ribonuclease A- and B- chain insulin. The N-terminal residues of the peptide fragments liberated were identified. The enzyme attacked both substrates only at alanyl, aspartyl, and asparaginyl residues, releasing these as free amino acids. In addition to alanine, aspartic acid, and asparagine, certain other amino acids were liberated from ribonuclease A, but these were accounted for by the relation of their position to alanine, aspartic acid, and asparagine residues.  相似文献   

13.
Deamidation of asparagine and glutamine residues, isomerization of aspartic acid side chains, and racemization of the L- to the D-form of the amino acids are common spontaneous chemical reactions known to occur in proteins. Previous studies have implicated succinimides as intermediates in these reactions; however, the evidence has been indirect. Our results demonstrate, for the first time, the presence of a succinimide intermediate in an intact protein. The succinimide (cyclic imide) variant was isolated from thermally stressed recombinant methionyl human growth hormone (hGH) by high performance anion-exchange chromatography, further purified by reversed-phase high performance liquid chromatography, and analyzed by tryptic mapping. A later eluting tryptic peptide, compared with the native T12 peptide (residues 128-134, Leu-Glu-Asp-Gly-Ser-Pro-Arg), was analyzed by mass spectrometry (MS). This variant had a protonated molecular mass of 755.3 atomic mass units (u), as compared with 773.3 u for the native T12 peptide. A difference of 18 u, a loss of water, is consistent with the formation of a succinimide intermediate at Asp-130 of methionyl hGH. MS/MS analysis of the cyclic imide-containing peptide verified that the modification occurred at Asp-130. A difference of 18 u was also observed for the intact cyclic imide methionyl hGH variant (22,238 u), as measured by electrospray mass spectrometry, compared with native methionyl hGH (22,256 u).  相似文献   

14.
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   

15.
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   

16.
Isomerization and peptide bond cleavage at aspartic residue (Asp) in peptide models have been reported. In this study, the mechanisms and energies concerning the isomerization and peptide bond cleavage at Asp residue were investigated by the density functional theory (DFT) at B3LYP/6-311++G(d,p). The integral equation formalism-polarizable continuum model (IEF-PCM) was utilized to calculate solvation effect by single-point calculation of the gas-phase B3LYP/6-311++G(d,p)-optimized structure. Mechanisms and energies of the dehydration in isomerization reaction of Asp residue were comparatively analyzed with the deamidation reaction of Asn residue. The results show that the succinimide intermediate was formed preferentially through the step-wise reaction via the tetrahedral intermediate. The cleavage at C-terminus is more preferential than those at N-terminus. In comparison to isomerization, peptide bond cleavage is ~20 kcal mol?1 and lower in activation barrier than the isomerization. So, in this case, the isomerization of Asp is inhibited by the peptide bond cleavage.  相似文献   

17.
Deamidation of asparagine residues, a post-translational modification observed in proteins, is a common degradation pathway in monoclonal antibodies (mAbs). The kinetics of deamidation is influenced by primary sequence as well as secondary and tertiary folding. Analytical hydrophobic interaction chromatography (HIC) is used to evaluate hydrophobicity of candidate mAbs and uncover post-translational modifications. Using HIC, we discovered atypical heterogeneity in a highly hydrophobic molecule (mAb-1). Characterization of the different HIC fractions using LC/MS/MS revealed a stable succinimide intermediate species localized to an asparagine-glycine motif in the heavy chain binding region. The succinimide intermediate was stable in vitro at pH 7 and below and increased on storage at 25°C and 40°C. Biacore evaluation showed a decrease in binding affinity of the succinimide intermediate compared with the native asparagine molecule. In vivo studies of mAb-1 recovered from a pharmacokinetic study in cynomolgus monkeys revealed an unstable succinimide species and rapid conversion to aspartic/iso-aspartic acid. Mutation from asparagine to aspartic acid led to little loss in affinity. This study illustrates the importance of evaluating modifications of therapeutic mAbs both in vitro and in serum, the intended environment of the molecule. Potential mechanisms that stabilize the succinimide intermediate in vitro are discussed.  相似文献   

18.
J Momand  S Clarke 《Biochemistry》1987,26(24):7798-7805
We have been interested in the metabolic fate of proteins containing aspartyl succinimide (Asu) residues. These residues can be derived from the spontaneous rearrangement of Asp and Asn residues and from the spontaneous demethylation of enzymatically methylated L-isoAsp and D-Asp residues. Incubation of the synthetic hexapeptide N-Ac-Val-Tyr-Pro-Asu-Gly-Ala with the cytosolic fraction of human erythrocytes resulted in rapid cleavage of the prolyl-aspartyl succinimide bond producing the tripeptide N-Ac-Val-Tyr-Pro. The rate of this reaction is equal for both L- and D-Asu-containing peptides and is 10-fold greater than the rate of cleavage of a corresponding peptide containing a normal Pro-Asp linkage. When the aspartyl succinimide ring was replaced with an isoaspartyl residue, the cleavage rate was about 5 times that of the normal Pro-Asp peptide. The tripeptide-producing activity copurified on DEAE-cellulose chromatography with an activity that cleaves N-carbobenzoxy-Gly-Pro-4-methylcoumarin-7-amide, a post-proline endopeptidase substrate. These two activities were both inhibited by an antiserum to rat brain post-proline endopeptidase, and it appears that they are catalyzed by the same enzyme. This enzyme has a molecular weight of approximately 80,000 and is covalently labeled and inhibited by [3H]diisopropyl fluorophosphate. The facile cleavage of the succinimide- and isoaspartyl-containing peptides by this post-proline endopeptidase suggests that it may play a role in the metabolism of peptides containing altered aspartyl residues.  相似文献   

19.
Recent studies have suggested that the isomerization/racemization of aspartate residues in proteins increases in aged tissues. One such residue is Asp151 in lens‐specific αA‐crystallin. Although many isomerization/racemization sites have been reported in various proteins, the factors that lead to those modifications in proteins in vivo remain obscure. Therefore, an in vitro system is needed to assess the mechanisms of modifications of Asp under various conditions. Deamidation of Asn to Asp in proteins occurs more rapidly than isomerization/racemization of Asp, although the reaction passes through the same intermediate in both pathways. Here, therefore, we replaced Asp151 in human lens αA‐crystallin with Asn by using site‐directed mutagenesis. The recombinant protein was expressed in Escherichia coli and used to investigate the deamidation/isomerization/racemization of Asn151 after incubation at 50°C for various durations and under different pH. After incubation, the mutant αA‐crystallin was subjected to enzymatic digestion followed by liquid chromatography–MS/MS to evaluate the ratio of modifications in Asn151‐containing peptides. The Asp151Asn αA‐crystallin mutant showed rapid deamidation to Asp with the formation of specific Asp isomers. In particular, deamidation increased greatly under basic conditions. By contrast, subunit–subunit interactions between αA‐crystallin and αB‐crystallin had little effect on the modification of Asn151. Our findings suggest that the Asp151Asn αA‐crystallin mutant represents a good in vitro model protein to assess deamidation, isomerization, and the racemization intermediates. Furthermore, our in vitro results show a different trend from in vivo data, implying the presence of specific factors that induce racemization from L‐Asp to D‐Asp residues in vivo.  相似文献   

20.
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   

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