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.
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.
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   

4.
B A Johnson  D W Aswad 《Biochemistry》1985,24(10):2581-2586
At pH 7.4, 37 degrees C, bovine brain protein carboxyl methyltransferase transiently methylates deamidated adrenocorticotropin. The methylation occurs at the alpha-carboxyl group of an atypical beta-carboxyl-linked isoaspartyl residue (position 25). Several lines of evidence indicate that the immediate product of demethylation is an aspartyl cyclic imide involving positions 25 and 26. The evidence includes (1) the rapid rate of methyl ester hydrolysis, which is consistent with intramolecular catalysis, (2) the inability of the demethylated product to be remethylated, (3) the charge of this product, and (4) its rate of breakdown. The eventual hydrolysis of the cyclic imide produces a 30/70 mixture of peptides containing either alpha- or beta-carboxyl-linked aspartyl residues, respectively. Cyclic imide formation is nonenzymatic and can explain the unusual lability of mammalian protein methyl esters in general. These findings suggest that protein carboxyl methylation in mammalian tissues is not a simple on/off reversible modification as it apparently is in chemotactic bacteria. Carboxyl methylation may serve to activate selected protein carboxyl groups for subsequent longer lasting modifications, possibly subserving a role in protein repair, degradation, cross-linking, or some other as yet undiscovered alteration of protein structure.  相似文献   

5.
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   

6.
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   

7.
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.
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   

10.
Jakas A  Horvat S 《Biopolymers》2003,69(4):421-431
Reactions between biological amines and reducing sugars (the Maillard reaction) are among the most important of the chemical and oxidative changes occurring in biological systems that contribute to the formation of a complex family of rearranged and dehydrated covalent adducts that have been implicated in the pathogenesis of human diseases. In this study, chemistry of the Maillard reactions was studied in four model systems containing fructosamines (Amadori compounds) obtained from the endogenous opioid pentapeptide leucine-enkephalin (Tyr-Gly-Gly-Phe-Leu), leucine-enkephalin methyl ester, structurally related tripeptide (Tyr-Gly-Gly), or from amino acid (Tyr). The degradation of model compounds as well as their ability to develop Maillard fluorescence was investigated under oxidative conditions in methanol and phosphate buffer pH 7.4 at two different temperatures (37 and 70 degrees C). At 37 degrees C, glycated leucine-enkephalin degraded slowly in methanol (t(1/2) approximately 13 days) and phosphate buffer (t(1/2) approximately 9 days), producing a parent peptide compound as a major product throughout a three-week incubation period. Whereas fluorescence slowly increased over time at 37 degrees C, incubations off all studied Amadori compounds at 70 degrees C resulted in a rapid appearance of a brown color and sharp increase in AGE (advanced glycation end products)-associated fluorescence (excitation 320 nm/emmision 420 nm) as well as in distinctly higher amounts of fragmentation products. The obtained data indicated that the shorter the peptide chain the more degradation products were formed. These studies have also helped to identify a new chemical transformation of the peptide backbone in the Maillard reaction that lead to beta-scission of N-terminal tyrosine side chain and p-hydroxybenzaldehyde formation under both aqueous and nonaqueous conditions.  相似文献   

11.
Protein L-isoaspartyl methyltransferase provides a basis for enzymatic measurement of atypical, isoaspartyl linkages which make a major contribution to protein microheterogeneity. The low Vmax of the methyltransferase reaction and the instability of the methyl ester can hinder accurate determinations, and different laboratories using different conditions have achieved discrepant values for the isoaspartate content of the same proteins. To investigate the effects of these conditions, and to optimize the assay, isoaspartyl delta sleep-inducing peptide was methylated under a variety of conditions. We found that 1 microM methyltransferase was required to obtain stoichiometric modification of 2 microM peptide in 40-min reactions at pH 6.2 and 30 degrees C. A computer model utilizing kinetic constants obtained from studies on initial rates of methylation predicted the same requirement for enzyme concentration. Carrier protein was necessary for optimal methyltransferase activity at enzyme concentrations below 0.4 microM. Stoichiometric methylation required concentrations of S-adenosylmethionine to be in substantial excess over those of peptide; 50 microM S-adenosylmethionine is the minimum needed for complete modification of 10 microM peptide. Spontaneous demethylation was significant under all conditions tested, so that the methyl ester itself never reached a ratio of 1 mol/mol of total peptide. These results demonstrate that the most accurate measurements of isoaspartate are obtained when reactions are carried out at low peptide concentrations, high S-adenosylmethionine concentrations, and high enzyme concentrations.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
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   

13.
New intravenous conjugates of amphotericin B (AMB) with poly(ethylene glycols) (PEG) (M=5000, 10,000, 20,000) have been synthesized and characterised. The intermediate PEGs possess a 1,4-disubstituted benzene ring with aldehyde group at the end of the chain. The benzene ring is connected with PEG at its 4-position (with respect to the aldehyde group) by various functional groups (ether, amide, ester). Reaction of terminal aldehyde group of the substituted PEGs with AMB gave conjugates containing a pH-sensitive imine linkage, which can be presumed to exhibit antimycotic effect at sites with lowered pH value. All types of the conjugates are relatively stable in phosphate buffer at physiological conditions of pH 7.4 (37 degrees C), less than 5 mol% AMB being split off from them within 24 h. For a model medium of afflicted tissue was used a phosphate buffer (pH 5.5, 37 degrees C), in which controlled release of AMB from the conjugates takes place. The imine linkage is split to give free AMB with half-lives of 2-45 min. The rate of acid catalysed hydrolysis depends upon substitution of the benzene ring; however, it does not depend on molecular weights of the PEGs used. The conjugates with ester linkage undergo enzymatic splitting in human blood plasma and/or blood serum at pH 7.4 (37 degrees C) with half-lives of 2-5 h depending on molecular weights of the PEGs used (M = 5000, 10,000, 20,000). At first, the splitting of ester linkage produces the relatively stable pro-drug, that is, 4-carboxybenzylideniminoamphotericin B, which is decomposed to AMB and 4-formylbenzoic acid in a goal-directed manner only at pH 7 (t1/2 = 2 min, pH 5.5, 37 degrees C). A goal-directed release of AMB is only achieved by acid catalysed hydrolysis of imine linkage, either from the polymeric conjugate or from the pro-drug released thereof. The LD50 values determined in vivo (mouse) are 20.7 mg/kg and 40.5 mg/kg for the conjugates with ester linkage (M = 10,000 and 5000, respectively), which means that they are ca. 6-11 times less toxic than free AMB.  相似文献   

14.
Selective deamidation of proteins and peptides is a reaction of great interest, both because it has a physiological role and because it can cause alteration in the biological activity, local folding, and overall stability of the protein. In order to evaluate the thermodynamic effects of this reaction in proteins, we investigated the temperature-induced denaturation of ribonuclease A derivatives in which asparagine 67 was selectively replaced by an aspartyl residue or an isoaspartyl residue, as a consequence of an in vitro deamidation reaction. Differential scanning calorimetry measurements were performed in the pH range 3.0-6.0, where the unfolding process is reversible, according to the reheating criterion used. It resulted that the monodeamidated forms have a different thermal stability with respect to the parent enzyme. In particular, the replacement of asparagine 67 with an isoaspartyl residue leads to a decrease of 6.3 degrees C of denaturation temperature and 65 kJ mol-1 of denaturation enthalpy at pH 5.0. These results are discussed and correlated to the X-ray three-dimensional structure of this derivative. The analysis leads to the conclusion that the difference in thermal stability between RNase A and (N67isoD)RNase A is due to enthalpic effects arising from the loss of two important hydrogen bonds in the loop containing residue 67, partially counterbalanced by entropic effects. Finally, the influence of cytidine-2'-monophosphate on the stability of the three ribonucleases at pH 5.0 is studied and explained in terms of its binding on the active site of ribonucleases. The analysis makes it possible to estimate the apparent binding constant and binding enthalpy for the three proteins.  相似文献   

15.
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   

16.
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   

17.
The plasma proteins, alpha 2-macroglobulin and complement components 3 and 4, contain an internal thiol ester involving a glutamyl and cysteinyl residue. The thiol ester is susceptible to cyclization at greater than 37 degrees C and forms an unstable 5-oxyproline intermediate. The latter can be hydrolyzed to produce two peptide fragments. We propose that enzymes having activated glutamyl residues as part of their catalytic mechanisms may undergo an analogous cyclization and peptidyl cleavage. As a model, we have investigated pig heart succinyl-CoA:3-keto acid transferase. When the CoA-enzyme thiolester intermediate is heated at pH 7.4 and 70 degrees C for 1 h, approximately 60% of the Mr = 60,000 subunits are cleaved to give Mr = 40,000 and 20,000 fragments. We have shown that formation of the enzyme thiolester is an obligate precursor for the protein fragmentation. However, the reaction was incomplete with a maximum of approximately 65% cleavage at times greater than 60 min. These results suggest that there is a competing, deactivation reaction; namely, the thiol ester and oxyproline intermediates are hydrolyzed to regenerate the active site glutamic acid. Although the maximum rate of cleavage is at 70 degrees C, approximately 15% autolysis also occurs at 37 degrees C. The Mr = 40,000 fragment had the same amino terminal sequence as the Mr = 60,000 subunit, (Trp-Lys-Phe-Tyr-Thr-Asp-Ala-Val-Glu-Ala-). No amino terminal could be detected for the Mr = 20,000 fragment, even after digesting the fragment with pyroglutaminase. Peptide maps of the fragments and the uncleaved subunit indicate that the fragments are generated in parallel. The size of the fragments puts the active site about two-thirds of the way from the amino terminal of the protein.  相似文献   

18.
Human erythrocyte membranes (ghosts) prepared from fresh blood changed in shape from spherical to crenated, when suspended in 10(-7)-10(-6) M Ca2+-EGTA buffers. Although the ghosts from long-stored ACD blood (10 weeks) were less sensitive to 10(-7)-10(-6) M Ca2+, the ghosts obtained from this blood after it had been preincubated with adenine and inosine for 3 h at 37 degrees C were highly sensitive to Ca2+. When these highly sensitive ghosts were incubated in 10 mM Tris-Cl buffer (pH 7.4) or 1 mM MgCl2 (pH 7.4) at 0 degrees C, they gradually lost Ca2+ sensitivity within 60 min, but they recovered Ca2+ sensitivity again after re-incubation with 2 mM Mg-ATP for 20 min at 37 degrees C followed by washing with 1 mM MgCl2 (pH 7.4). The shape of these highly Ca2+-sensitive ghosts immediately changed from crenate to disc on addition of 1 mM Mg-ATP even at 6 degrees C in the presence of 10(-7)-10(-6) M Ca2+. A similar shape change was also observed when ghosts treated with 0.5% Triton X-100 (Triton shells) were used. Triton shells from fresh blood ghosts or from long-stored blood ghosts which had been preincubated with 2 mM Mg-ATP for 20 min at 37 degrees C shrank immediately in the presence of 10(-6) M Ca2+ and then swelled on addition of 1 mM Mg-ATP. The specificity to ATP and the dependency on ATP concentration are in agreement with those of the ghost shape change at step 2 (Jinbu, Y. et al., Biochem biophys res commun 112 (1983) 384-390) [18]. These results suggest that cytoskeletal protein phosphorylation enhances sensitivity to Ca2+ and induces erythrocyte shape change in the presence of physiological concentrations of ATP and Ca2+.  相似文献   

19.
The intermediate produced from dCyd by HNO2 and NO treatments was isolated and characterized. When 10 mM dCyd was treated with 100 mM NaNO2 in 1.0 M acetate buffer (pH 3.7) at 37 degrees C, a previously unidentified product was formed. By spectrometric measurements, the product was identified as a diazoate derivative of dCyd, 1-(beta-D-2'-deoxyribofuranosyl)-2-oxopyrimidine-4-diazoate. The time course of the concentration change of the diazoate showed a profile characteristic of a reaction intermediate, and the maximum yield was 37 microM at the reaction time of 25 min. Up to the reaction time of 10 min, the diazoate concentration was greater than that of dUrd, a deamination product of dCyd. Addition of thiocyanate increased the yield of the diazoate in HNO2 treatment, whereas addition of ascorbate decreased the yield. When 10 mM dCyd in 100 mM phosphate buffer was treated with NO at 37 degrees C under aerobic conditions holding the pH (7.2-7.6), the diazoate was also generated. The yield of the diazoate was higher than that of dUrd up to 15 mmol of NO absorption. At pH 3.7 and 37 degrees C, the diazoate was converted to dUrd with the first-order rate constant k = 4.8 x 10(-)4 s-1 (t1/2 = 24 min). Under physiological conditions (pH 7.4, 37 degrees C), however, it was fairly stable (k = 5.8 x 10(-)7 s-1, t1/2 = 330 h). In both cases, the diazoate was converted to dUrd exclusively and no other intermediates were detected by HPLC analysis. Uracil-DNA glycosylase did not remove the diazoate residue from an oligodeoxynucleotide containing this damage, [d(T6DT5), D = the diazoate]. The Tm value of a duplex containing the diazoate, d(T6DT5).d(A5GA6), was much lower than that of a duplex containing a correct C:G base pair, d(T6CT5).d(A5GA6). These results show that the diazoate is generated as a stable intermediate in the reactions of dCyd with HNO2 and NO and that the major product is the diazoate but not dUrd in the initial stage of the reactions. Thus, once formed in vivo, the diazoate persists for long time in DNA and may act as a major cytotoxic and/or genotoxic lesion with biologically relevant doses of HNO2 and NO.  相似文献   

20.
The effect of delta-sleep-inducing peptide (DSIP) on erythrocytic membranes of human donor blood was studied by the spin label and spin probe methods. The spin-labeled derivative of DSIP containing the N-terminal residue of 1-oxyl-2,2,5,5-tetramethylpyrroline-3-carboxylic acid was synthesized. An analysis of the ESR spectra of the spin-labeled DSIP derivative recorded after its incubation with a human erythrocyte suspension at 37 degrees C revealed a decrease in the rotational correlation time (tau c) and molecular order parameter (S) in comparison with the control solutions of the peptide in phosphate buffer (pH 7.4). The application of paramagnetic probes, 5-, 12-, and 16-doxylstearic acids and 3-doxylandrostanol, demonstrated that the introduction of DSIP in an erythrocytic suspension significantly increased the mobility of the hydrophobic area of the membrane bilayer both at a depth of 20-22 A and in the subsurface area (4-6 A). The dependence of these effects on the DSIP concentration was shown to have the form of a curve with well-defined extremes. The maximal disordering of membrane lipids was observed at peptide concentrations of 10(-9) and 10(-6) M. These results suggested that DSIP significantly affected the structure of plasmatic membranes in vitro by changing the physical state of their lipid components.  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