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.
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   

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.
The spontaneous degradation of asparaginyl and aspartyl residues to isoaspartyl residues is a common type of protein damage in aging organisms. Although the protein-l-isoaspartyl (d-aspartyl) O-methyltransferase (EC 2.1.1.77) can initiate the repair of l-isoaspartyl residues to l-aspartyl residues in most organisms, no gene homolog or enzymatic activity is present in the budding yeast Saccharomyces cerevisiae. Therefore, we used biochemical approaches to elucidate how proteins containing isoaspartyl residues are metabolized in this organism. Surprisingly, the level of isoaspartyl residues in yeast proteins (50–300 pmol of isoaspartyl residues/mg of protein extract) is comparable with organisms with protein-l-isoaspartyl (d-aspartyl) O-methyltransferase, suggesting a novel regulatory pathway. Interfering with common protein quality control mechanisms by mutating and inhibiting the proteasomal and autophagic pathways in vivo did not increase isoaspartyl residue levels compared with wild type or uninhibited cells. However, the inhibition of metalloproteases in in vitro aging experiments by EDTA resulted in an ∼3-fold increase in the level of isoaspartyl-containing peptides. Characterization by mass spectrometry of these peptides identified several proteins involved in metabolism as targets of isoaspartyl damage. Further analysis of these peptides revealed that many have an N-terminal isoaspartyl site and originate from proteins with short half-lives. These results suggest that one or more metalloproteases participate in limiting isoaspartyl formation by robust proteolysis.  相似文献   

4.
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   

5.
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   

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.
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   

8.
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   

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.
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   

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.
We have found that a chicken egg lysozyme derivative (beta-101-lysozyme) containing an L-isoaspartyl residue at position 101 has a Km for methylation by the human erythrocyte L-isoaspartyl/D-aspartyl protein methyltransferase (EC 2.1.1.77) of 183 microM, about 30 times higher than that expected from previous studies with isoaspartyl-containing peptides. In the course of investigating the reasons for this poor enzyme recognition, we found that charged residues on the carboxyl side of isoaspartyl residues had a large effect on the affinity of the enzyme for synthetic peptides. This is best illustrated by the lysozyme-related peptide YVSisoDGDG, which has a Km for methylation of 469 microM. When the penultimate aspartyl residue is replaced by a cysteinyl residue, the Km drops to 4.6 microM, comparable to other peptides of similar size. Furthermore, replacing it with a cysteic acid residue results in a Km of 104 microM, suggesting that a negative charge at this position may lead to a weaker affinity of the peptide substrate for the methyltransferase. Assays with additional synthetic peptides indicate that moving the negative charge to the first or third residue on the carboxyl side of the isoaspartyl residue has a similar but less severe effect in reducing its affinity for the methyltransferase. Enzymatic methylation has recently been proposed to be the first step in the conversion of abnormal isoaspartyl residues to aspartyl residues. The results reported here, however, along with previous evidence that protein tertiary structure can inhibit isoaspartyl methylation, suggest that only a subclass of damaged sites are capable of efficiently entering a putative repair pathway; the sites not recognized by the methyltransferase may accumulate in vivo.  相似文献   

13.
The formation of isoaspartate (isoAsp) from asparaginyl or aspartyl residues is a spontaneous post-translational modification of peptides and proteins. Due to isopeptide bond formation, the structure and possibly function of peptides and proteins is altered. IsoAsp modifications within the peptide chain have been reported for many cytosolic proteins. Amyloid peptides (Aβ) deposited in Alzheimer’s disease may carry an N-terminal isoAsp-modification. Here, we describe a quantitative investigation of isoAsp-formation from N-terminal Asn and Asp using model peptides similar to the Aβ N-terminus. The study is based on a newly developed separation of peptides using capillary electrophoresis (CE). 1H NMR was employed to validate the basic finding of N-terminal isoAsp-formation from Asp and Asn. Thereby, the isomerization of Asn at neutral pH (0.6 day?1, peptide NGEF) is approximately six times faster than that within the peptide chain (AANGEF). The difference in velocity between Asn and Asp isomerization is approximately 50-fold. In contrast to N-terminal Asn, Asp isomerization is significantly accelerated at acidic pH. The kinetic solvent isotope (k D2O/k H2O) effect of 2.46 suggests a rate-limiting proton transfer in isoAsp-formation. The proton inventory is consistent with transfer of one proton in the transition state, supporting the previous notion of rate-limiting deprotonation of the peptide backbone amide during succinimide-intermediate formation. The study provides evidence for a spontaneous N-terminal isoAsp-formation within peptides and might explain the accumulation of N-terminal isoAsp in amyloid deposits.  相似文献   

14.
Amyloid fibril formation is widely accepted as a critical step in all types of amyloidosis. Amyloid fibrils derived from different amyloidogenic proteins share structural elements including beta-sheet secondary structure and similar tertiary structure. While some amyloidogenic proteins are rich in beta-sheet in their soluble form, others, like Alzheimer beta-amyloid peptide (Abeta) or serum amyloid A, must undergo significant structural transition to acquire a high beta-sheet content. We postulate that Abeta and other amyloidogenic proteins undergo a transition to beta-sheet as a result of aging-related chemical modifications of aspartyl residues to the form of succinimide or isoaspartyl methyl ester. We hypothesize that spontaneous cyclization of aspartate residues in amyloidogenic proteins can serve as a nucleation event in amyloidogenesis. To test this hypothesis, we synthesized a series of designed peptides having the sequence VTVKVXAVKVTV, where X represents aspartic acid or its derivatives. Studies using circular dichroism showed that neutralization of the aspartate residue through the formation of a methyl ester or an amide, or replacement of aspartate with glutamate led to an increased beta-sheet content at neutral and basic pH. A higher content of beta-sheet structure correlated with increased propensity for fibril formation and decreased solubility at neutral pH.  相似文献   

15.
A simple and site-specific nonenzymatic method based on pyrolysis has been developed to cleave peptides and proteins. Pyrolytic cleavage was found to be specific and rapid as it induced a cleavage at the C-terminal side of aspartic acid in the temperature range of 220-250 degrees C in 10 s. Electrospray ionization (ESI) mass spectrometry (MS) and tandem-MS (MS/MS) were used to characterize and identify pyrolysis cleavage products, confirming that sequence information is conserved after the pyrolysis process in both peptides and protein tested. This suggests that pyrolysis-induced cleavage at aspartyl residues can be used as a rapid protein digestion procedure for the generation of sequence-specific protein biomarkers.  相似文献   

16.
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   

17.
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   

18.
The isomerization rate of aspartic acid (Asp) residue is known to be affected by the three-dimensional structures of peptides and proteins. Although the isomerized Asp residues were experimentally observed, structural features which affect the isomerization cannot be elucidated sufficiently because of protein denaturation and aggregation. In this study, molecular dynamics (MD) simulations were conducted on three αA-crystallin peptides (T6, T10, and T18), each containing a single Asp residue with different isomerization rate (T18 > T6 > T10) to clarify the structural factors of Asp isomerization tendency. For MD trajectories, distances between side-chain carboxyl carbon of Asp and main-chain amide nitrogen of (n + 1) residue (Cγ–N distances), root mean square fluctuations (RMSFs), and polar surface areas for main-chain amide nitrogen of (n + 1) residues (PSAN) were calculated, because these structural features are considered to relate to the formations of cyclic imide intermediates. RMSFs and PSAN are indexes of peptide backbone flexibilities and solvent exposure of the amide nitrogen, respectively. The average Cγ–N distances of T10 was longer than those of the other two peptides. In addition, the peptide containing Asp residue with a higher isomerization rate showed higher flexibility of the peptide backbone around the Asp residue. PSAN for amide nitrogen in T18 were much larger than those of other two peptides. The computational results suggest that Asp-residue isomerization rates are affected by these factors.  相似文献   

19.
In prion-related encephalopathies, the cellular prion protein (PrP(C)) undergoes a change in conformation to become the scrapie prion protein (PrP(Sc)) which forms infectious deposits in the brain. Conceivably, the conformational transition of PrP(C) to PrP(Sc) might be linked with posttranslational alterations in the covalent structure of a fraction of the PrP molecules. We tested a synthetic peptide corresponding to residues 106-126 of human PrP for the occurrence of spontaneous chemical modifications. The only asparagine residue, Asn108, was deamidated to aspartic acid and isoaspartic acid with a half-life of about 12 days. The same posttranslational modifications were found in recombinant murine full-length protein. On aging, 0.8 mol of isoaspartyl residue per mole of protein was detected by the protein-l-isoaspartyl methyltransferase assay (t(1/2) approximately 30 days). Mass spectrometry and Edman degradation of Lys-C fragments identified Asn108 in the amino-terminal flexible part of the protein to be partially converted to aspartic acid and isoaspartic acid. A second modification was the partial isomerization of Asp226' which is only present in rodents.  相似文献   

20.
l-Isoaspartyl (d-aspartyl) O-methyltransferase (PCMT1) can initiate the conversion of damaged aspartyl and asparaginyl residues to normal l-aspartyl residues. Mice lacking this enzyme (Pcmt1-/- mice) have elevated levels of damaged residues and die at a mean age of 42 days from massive tonic-clonic seizures. To extend the lives of the knockout mice so that the long term effects of damaged residue accumulation could be investigated, we produced transgenic mice with a mouse Pcmt1 cDNA under the control of a neuron-specific promoter. Pcmt1 transgenic mice that were homozygous for the endogenous Pcmt1 knockout mutation ("transgenic Pcmt1-/- mice") had brain PCMT1 activity levels that were 6.5-13% those of wild-type mice but had little or no activity in other tissues. The transgenic Pcmt1-/- mice lived, on average, 5-fold longer than nontransgenic Pcmt1-/- mice and accumulated only half as many damaged aspartyl residues in their brain proteins. The concentration of damaged residues in heart, testis, and brain proteins in transgenic Pcmt1-/- mice initially increased with age but unexpectedly reached a plateau by 100 days of age. Urine from Pcmt1-/- mice contained increased amounts of peptides with damaged aspartyl residues, apparently enough to account for proteins that were not repaired intracellularly. In the absence of PCMT1, proteolysis may limit the intracellular accumulation of damaged proteins but less efficiently than in wild-type mice having PCMT1-mediated repair.  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