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.
Under favorable conditions, Asp or Asn residues can undergo rearrangement to a succinimide (cyclic imide), which may also serve as an intermediate for deamidation and/or isoaspartate formation. Direct identification of such succinimides by peptide mapping is hampered by their lability at neutral and alkaline pH. We determined that incubation in 2 M hydroxylamine, 0.2 M Tris buffer, pH 9, for 2 h at 45 degrees C will specifically cleave on the C-terminal side of succinimides without cleavage at Asn-Gly bonds; yields are typically approximately 50%. N-terminal sequence analysis can then be used to identify an internal sequence generated by cleavage of the succinimide, hence identifying the succinimide site.  相似文献   

2.
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   

3.
We examined the effects of osmolytes, sucrose and trehalose, on the deterioration of hen lysozyme as a model protein. Sucrose and trehalose depressed the aggregation of lysozyme molecules caused by heating at 100 degrees C at pH 6. Since lysozyme was fully denatured under these conditions, the effects of sucrose and trehalose on the denatured state of lysozyme were investigated using reduced S-alkylated lysozyme, a model of denatured hen lysozyme. From analyses of circular dichroism spectra and fluorescence spectra, sucrose and trehalose were found to induce alpha-helical conformations and some tertiary structures around tryptophan residues in the reduced S-alkylated lysozyme. Moreover, these compounds also depressed chemical reactions such as deamidation and racemization, which often cause the deterioration of proteins, on the reduced S-alkylated lysozyme. Therefore, the data suggest that sucrose and trehalose have a propensity to depress such deterioration as the aggregation of protein molecules or chemical reactions in proteins by inducing some tertiary structures (including alpha-helical structures) in the polypeptide chain.  相似文献   

4.
Protein cyclic imide is the putative intermediate in the formation of sites of carboxyl-methylation in eukaryotic proteins. Conditions known to induce the formation of a cyclic imide in model peptides have been applied to a protein, calmodulin. Heating of calmodulin in the dry state at 100 degrees C for 24 h after lyophilization from a pH 2.0 or pH 6.0 solution produces derivatives with altered chromatographic properties in anion-exchange HPLC. At pH 6.0, complete activity of calmodulin was retained. Analysis with Fourier transform infrared (FTIR)-photoacoustic spectroscopy demonstrated the presence of a new structure in the calmodulin molecule consistent with modification of carboxylic acid groups. The conversion of calmodulin is dependent upon the absence of Ca2+ (the presence of 1 mM ethylene glycol bis(beta-aminoethyl ether) N,N'-tetraacetic acid). A peptide analogous to the calcium binding regions of calmodulin, Asp-Lys-Asp-Gly-Asn-Gly-Thr-Ile-Thr-Thr-Lys-Glu, is also converted, upon heating, to chromatographically different forms in reversed-phase chromatography. This process is also dependent upon the absence of calcium. Sequence analysis of the peptide derivatives reveals a second amino terminus, implicating peptide bond hydrolysis in the product. A dipeptide, Asp-Gly, known to form a cyclic imide structure under similar conditions is also hydrolyzed during sequence analysis consistent with cleavage occurring at the position of the cyclic imide structure. Asp3 is suggested to be the site of cyclic imide formation in the calmodulin peptide. The presence of a cyclic imide structure is also confirmed by the application of FTIR-photoacoustic spectroscopy. These data suggest that cyclic imide formation in calmodulin has been induced, possibly at one, or more, of the calcium binding loops of the protein. These modification reactions may provide a basis for future investigations of cyclic imide formation in proteins.  相似文献   

5.
In contrast with the common belief that all the amino acid residues in higher organisms are l ‐forms, d ‐amino acid residues have been recently detected in various aging tissues. Aspartic acid (Asp) residues are known to be the most prone to stereoinvert via cyclic imide intermediate. Although the glutamic acid (Glu) is similar in chemical structure to Asp, little has been reported to detect d ‐Glu residues in human proteins. In this study, we investigated the mechanism of the Glu‐residue stereoinversion catalyzed by water molecules using B3LYP/6‐31+G(d,p) density functional theory calculations. We propose that the Glu‐residue stereoinversion proceeds via a cyclic imide intermediate, i.e., glutarimide (GI). All calculations were performed by using a model compound in which a Glu residue was capped with acetyl and methylamino groups on the N‐ and C‐termini, respectively. We found that two water molecules catalyze the three steps involved in the GI formation: iminolization, cyclization, and dehydration. The activation energy required for the Glu residue to form a GI intermediate was estimated to be 32.3 kcal mol?1, which was higher than that of the experimental Asp‐residue stereoinversion. This calculation result suggests that the Glu‐residue stereoinversion is not favored under the physiological condition.  相似文献   

6.
Interpretation of deglycosylation studies relies heavily on the absence of modifications to the polypeptide chain. We have found that by using a common chemical deglycosylation technique, one can effect at least three changes in a peptide's structure: methylation, isomerization, and ring formation. It was determined that the conditions of chemical deglycosylation introduce a + 14 Da shift in the masses of our model peptides, RKDVY, RKEVY, and horseradish peroxidase. This shift is localized to acidic functional groups and is interpreted as methylation of the free carboxylates in our models. An additional shift in mass of – 18 Da is found in the model peptide RKDVY consistent with the loss of water associated with succinimide ring formation in this peptide. Chemical treatment induced isomerization of aspartyl residues to isoaspartyl residues in another model peptide, tetragastrin. These results indicate that one should use caution when interpreting the results of chemical deglycosylation experiments.  相似文献   

7.
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   

8.
Proteins are vulnerable to spontaneous, covalent modifications that may result in alterations to structure and function. Asparagines are particularly labile, able to undergo deamidation through the formation of a succinimide intermediate to produce either aspartate or isoaspartate residues. Although aspartates cannot undergo deamidation they can form a succinimide and result in the same products. Isoaspartyls are the principal product of succinimide hydrolysis, accounting for 65-85% of the emerging residues. The variability in the ratio of products emerging from succinimide hydrolysis suggests the ability of protein structure to influence succinimide outcome. In the H15D histidine-containing protein (HPr), phosphorylation of the active site aspartate catalyzes the formation of a cyclic intermediate. Resolution of this species is exclusively to aspartate residues, suggestive of either a succinimide with restrained hydrolysis, or an isoimide, from which aspartyl residues are the only possible product. Deletion of the C-terminal residue of this protein does not influence the ability for phosphorylation or ring formation, but it does allow for isoaspartyl formation, verifying a succinimide as the cyclic intermediate in H15D HPr. Isoaspartyl formation in H15D Delta85 is rationalized to occur as a consequence of elimination of steric restrictions imposed by the C terminus on the main-chain carbonyl of the succinimide, the required point of nucleophilic attack of a water molecule for isoaspartyl formation. This is the first reported demonstration of the influence of protein structure on the products emerging from succinimide hydrolysis.  相似文献   

9.
Aspartate129 in porcine somatotropin was converted into a cyclic imide residue (succinimide) under acidic solution conditions. Reversed-phase high performance liquid chromatography was utilized to isolate and quantitate this altered species, which accounted for approximately 30% of the total protein. The molecular mass of this modified species was determined by electrospray mass spectrometry to be 18 Da less than normal porcine somatotropin, indicative of a loss of 1 H2O molecule. Tryptic peptide mapping demonstrated that the peptide composed of residues 126-133 was altered in this modified protein. Amino acid analysis, amino acid sequencing, mass spectrometry, and capillary zone electrophoresis were used to demonstrate that aspartate129 in this peptide had been converted into a succinimide residue. Further confirmation that this peptide contained a succinimide was obtained by hydrolyzing the modified peptide at pH 9.0, which yielded both the aspartate and isoaspartate peptides.  相似文献   

10.
The objective of this study was to explain the increased propensity for the conversion of cyclo-(1,7)-Gly-Arg-Gly-Asp-Ser-Pro-Asp-Gly-OH (1), a vitronectin-selective inhibitor, to its cyclic imide counterpart cyclo-(1,7)-Gly-Arg-Gly-Asu-Ser-Pro-Asp-Gly-OH (2). Therefore, we present the conformational analysis of peptides 1 and 2 by NMR and molecular dynamic simulations (MD). Several different NMR experiments, including COSY, COSY-Relay, HOHAHA, NOESY, ROESY, DQF-COSY and HMQC, were used to: (a) identify each proton in the peptides; (b) determine the sequential assignments; (c) determine the cis-trans isomerization of X-Pro peptide bond; and (d) measure the NH-HCalpha coupling constants. NOE- or ROE-constraints were used in the MD simulations and energy minimizations to determine the preferred conformations of cyclic peptides 1 and 2. Both cyclic peptides 1 and 2 have a stable solution conformation; MD simulations suggest that cyclic peptide 1 has a distorted type I beta-turn at Arg2-Gly3-Asp4-Ser5 and cyclic peptide 2 has a pseudo-type I beta-turn at Ser5-Pro6-Asp7-Gly1. A shift in position of the type I beta-turn at Arg2-Gly3-Asp4-Ser5 in peptide 1 to Ser5-Pro6-Asp7-Gly1 in peptide 2 occurs upon formation of the cyclic imide at the Asp4 residue. Although the secondary structure of cyclic peptide 1 is not conducive to succinimide formation, the reaction proceeds via neighbouring group catalysis by the Ser5 side chain. This mechanism is also supported by the intramolecular hydrogen bond network between the hydroxyl side chain and the backbone nitrogen of Ser5. Based on these results, the stability of Asp-containing peptides cannot be predicted by conformational analysis alone; the influence of anchimeric assistance by surrounding residues must also be considered.  相似文献   

11.
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   

12.
The l-α-Asp residues in peptides or proteins are prone to undergo nonenzymatic reactions to form l-β-Asp, d-α-Asp, and d-β-Asp residues via a succinimide five-membered ring intermediate. From these three types of isomerized aspartic acid residues, particularly d-β-Asp has been widely detected in aging tissue. In this study, we computationally investigated the cyclization of α- and β-Asp residues to form succinimide with dihydrogen phosphate ion as a catalyst (H2PO4). We performed the study using B3LYP/6-31 + G(d,p) density functional theory calculations. The comparison of the activation barriers of both residues is discussed. All the calculations were performed using model compounds in which an α/β-Asp-Gly sequence is capped with acetyl and methylamino groups on the N- and C-termini, respectively. Moreover, H2PO4 catalyzes all the steps of the succinimide formation (cyclization-dehydration) acting as a proton-relay mediator. The calculated activation energy barriers for succinimide formation of α- and β-Asp residues are 26.9 and 26.0 kcal mol 1, respectively. Although it was experimentally confirmed that β-Asp has higher stability than α-Asp, there was no clear difference between the activation barriers. Therefore, the higher stability of β-Asp residue than α-Asp residue may be caused by an entropic effect associated with the succinimide formation.  相似文献   

13.
We have investigated the formation of D-aspartyl and L-isoaspartyl (beta-aspartyl) residues and their subsequent methylation in bovine brain calmodulin by the type II protein carboxyl methyltransferase. Based on the results of studies with unstructured peptides and denatured proteins, it has been proposed that the major sites of carboxyl methylation in calmodulin are at L-isoaspartyl residues that originate from two Asn-Gly sequences. To test this hypothesis, we directly identified the sites of methylation in affinity-purified preparations of calmodulin by peptide mapping using the proteases trypsin, endoproteinase Lys-C, clostripain, chymotrypsin, and Staphylococcus aureus V8 protease. We found, however, that the major high-affinity sites of methylation originate from aspartyl residues at position 2 and at positions 78 and/or 80. The methylatable residue in the first case was shown to be L-isoaspartate by comparison of the properties of a synthetic peptide corresponding to the N-terminal 13 residues substituted with an L-iso-Asp residue at position 2. The second methylatable residue, probably derived from Asp78, also appears to be an L-isoaspartyl residue. These sites appear to be readily accessible to the methyltransferase and are present in relatively flexible regions of calmodulin that may allow the spontaneous degradation reactions to occur that generate L-isoaspartyl residues via succinimide intermediates. Interestingly, the four calcium binding regions, each containing 3-4 aspartyl and asparaginyl residues (including the two Asn-Gly sequences), do not appear to contribute to the high-affinity methyl acceptor sites, even when calcium is removed prior to the methylation reaction. We propose that methylatable residues do not form at these sites because of the inflexibility of these regions when calcium is bound.  相似文献   

14.
In the biological proteins, aspartic acid (Asp) residues are prone to nonenzymatic isomerization via a succinimide (Suc) intermediate. Asp-residue isomerization causes the aggregation and the insolubilization of proteins, and is considered to be involved in various age-related diseases. Although Suc intermediate was considered to be formed by nucleophilic attack of the main-chain amide nitrogen of N-terminal side adjacent residue to the side-chain carboxyl carbon of Asp residue, previous studies have shown that the nucleophilic attack is more likely to proceed via iminol tautomer when the water molecules act as catalysts. However, the full pathway to Suc-intermediate formation has not been investigated, and the experimental analyses for the Asp-residue isomerization mechanism at atomic and molecular levels, such as the analysis of the transition state geometry, are difficult. In the present study, we computationally explored the full pathways for Suc-intermediate formation from Asp residues. The calculations were performed two types of reactant complexes, and all energy minima and TS geometries were optimized using B3LYP density functional methods. As a result, the SI-intermediate formation was divided into three processes, i.e., iminolization, cyclization, and dehydration processes, and the activation energies were calculated to be 26.1 or 28.4 kcal mol−1. These values reproduce the experimental data. The computational results show that abundant water molecules in living organisms are effective catalysts for the Asp-residue isomerization.  相似文献   

15.
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   

16.
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   

17.
Wolanin PM  Webre DJ  Stock JB 《Biochemistry》2003,42(47):14075-14082
Response regulator proteins are phosphorylated on a conserved aspartate to activate responses to environmental signals. An intrinsic autophosphatase activity limits the duration of the phosphorylated state. We have previously hypothesized that dephosphorylation might proceed through an intramolecular attack, leading to succinimide formation, and such an intramolecular dephosphorylation event is seen for CheY and OmpR during mass spectrometric analysis [Napper, S., Wolanin, P. M., Webre, D. J., Kindrachuk, J., Waygood, B., and Stock, J. B. (2003) FEBS Lett 538, 77-80]. Succinimide formation is usually associated with the spontaneous deamidation of Asn residues. We show here that an Asp57 to Asn mutant of the CheY chemotaxis response regulator undergoes an unusually rapid deamidation back to the wild-type Asp57, supporting the hypothesis that the active site of CheY is poised for succinimide formation. In contrast, we also show that the major route of phosphoaspartate hydrolysis in CheY occurs through water attack on the phosphorus both during autophosphatase activity and during CheZ-mediated dephosphorylation. Thus, CheY dephosphorylation does not usually proceed via a succinimide or any other intramolecular attack.  相似文献   

18.
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   

19.
Amyloid formation is associated with several human diseases including Alzheimer's disease (AD), Parkinson's disease, Type 2 Diabetes, and so forth, no disease modifying therapeutics are available for them. Because of the structural similarities between the amyloid species characterizing these diseases, (despite the lack of amino acid homology) it is believed that there might be a common mechanism of toxicity for these conditions. Thus, inhibition of amyloid formation could be a promising disease-modifying therapeutic strategy for them. Aromatic residues have been identified as crucial in formation and stabilization of amyloid structures. This finding was corroborated by high-resolution structural studies, theoretical analysis, and molecular dynamics simulations. Amongst the aromatic entities, tryptophan was found to possess the most amyloidogenic potential. We therefore postulate that targeting aromatic recognition interfaces by tryptophan could be a useful approach for inhibiting the formation of amyloids. Quinones are known as inhibitors of cellular metabolic pathways, to have anti- cancer, anti-viral and anti-bacterial properties and were shown to inhibit aggregation of several amyloidogenic proteins in vitro. We have previously described two quinone-tryptophan hybrids which are capable of inhibiting amyloid-beta, the protein associated with AD pathology, both in vitro and in vivo. Here we tested their generic properties and their ability to inhibit other amyloidogenic proteins including α-synuclein, islet amyloid polypeptide, lysozyme, calcitonin, and insulin. Both compounds showed efficient inhibition of all five proteins examined both by ThT fluorescence analysis and by electron microscope imaging. If verified in vivo, these small molecules could serve as leads for developing generic anti-amyloid drugs.  相似文献   

20.
The propensity to associate or aggregate is one of the characteristic properties of many nonnative proteins. The aggregation of proteins is responsible for a number of human diseases and is a significant problem in biotechnology. Despite this, little is currently known about the effect of self-association on the structural properties and conformational stability of partially folded protein molecules. G-actin is shown to form equilibrium unfolding intermediate in the vicinity of 1.5 M guanidinium chloride (GdmCl). Refolding from the GdmCl unfolded state is terminated at the stage of formation of the same intermediate state. An analogous form, known as inactivated actin, can be obtained by heat treatment, or at moderate urea concentration, or by the release of Ca(2+). In all cases actin forms specific associates comprising partially folded protein molecules. The structural properties and conformational stability of inactivated actin were studied over a wide range of protein concentrations, and it was established that the process of self-association is rather specific. We have also shown that inactivated actin, being denatured, is characterized by a relatively rigid microenvironment of aromatic residues and exhibits a considerable limitation in the internal mobility of tryptophans. This means that specific self-association can play an important structure-forming role for the partially folded protein molecules.  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