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.
Nature of the fast and slow refolding reactions of iron(III) cytochrome c   总被引:2,自引:0,他引:2  
The fast and slow refolding reactions of iron(III) cytochrome c (Fe(III) cyt c), previously studied by Ikai et al. (Ikai, A., Fish, W. W., & Tanford, C. (1973) J. Mol. Biol. 73, 165--184), have been reinvestigated. The fast reaction has the major amplitude (78%) and is 100-fold faster than the slow reaction in these conditions (pH 7.2, 25 degrees C, 1.75 M guanidine hydrochloride). We show here that native cyt c is the product formed in the fast reaction as well as in the slow reaction. Two probes have been used to test for formation of native cyt c. absorbance in the 695-nm band and rate of reduction of by L-ascorbate. Different unfolded species (UF, US) give rise to the fast and slow refolding reactions, as shown both by refolding assays at different times after unfolding ("double-jump" experiments) and by the formation of native cyt c in each of the fast and slow refolding reactions. Thus the fast refolding reaction is UF leads to N and the slow refolding reaction is Us leads to N, where N is native cyt c, and there is a US in equilibrium UF equilibrium in unfolded cyt c. The results are consistent with the UF in equilibrium US reaction being proline isomerization, but this has not yet been tested in detail. Folding intermediates have been detected in both reactions. In the UF leads to N reaction, the Soret absorbance change precedes the recovery of the native 695-nm band spectrum, showing that Soret absorbance monitors the formation of a folding intermediate. In the US leads to N reaction an ascorbate-reducible intermediate has been found at an early stage in folding and the Soret absorbance change occurs together with the change at 695 nm as N is formed in the final stage of folding.  相似文献   

2.
Pure [3H]ethyltubulin dimer, containing 1.07 mol of [3H]ethyl groups/110.000 g protein was prepared by reaction of tubulin with acetaldehyde and [3H]sodium borohydride. The derivatized tubulin dimer was shown to be native by the following criteria: (1) the stoichiometry for [3H]GDP binding was similar to that for native tubulin: (2) it repeatedly copolymerized and codepolymerized with native tubulin with constant specific activity. The potential utility for [3H]ethyltubulin in quantitating tubulin in biological samples by isotope dilution, and in studying the relationships between microtubules, rings, and dimers is discussed.  相似文献   

3.
Pyridoxamine is a vitamin B6 derivative involved in biological reactions such as transamination, and can also act as inhibitor in protein glycation. In both cases, it has been reported that Schiff base formation between pyridoxamine and carbonyl compounds is the main step. Nevertheless, few studies on the Schiff base formation have been reported to date. In this work, we conduct a comparative study of the reaction of pyridoxamine and 4-picolylamin (a pyridoxamine analog) with various carbonyl compounds including propanal, formaldehyde and pyruvic acid. Based on the results, 4-picolylamin forms a Schiff base as end-product of its reactions with propanal and pyruvic acid, but a carbinolamine with formaldehyde. On the other hand, pyridoxamine forms a Schiff base with the three reagents, but the end-product is in equilibrium with its hemiaminal form, which results from the attack of the phenolate ion of the pyridine ring on the imine carbon. This isomeric equilibrium should be considered in studying reactions involving amine derivatives of vitamin B6.  相似文献   

4.
C E Barry  P G Nayar  T P Begley 《Biochemistry》1989,28(15):6323-6333
Phenoxazinone synthase is a copper-containing oxidase that catalyzes the coupling of 2-aminophenols to form the 2-aminophenoxazinone chromophore. This reaction constitutes the final step in the biosynthesis of the potent antineoplastic agent actinomycin. The mechanism of this complex 6-electron oxidation was determined by using a variety of substituted 2-aminophenols, designed to block the reaction at intermediate stages. Thus, with 3,5-di-tert-butyl-2-aminophenol as substrate, the reaction was blocked at the o-quinone imine 17; with 5-tert-butyl-2-aminophenol (19) as substrate, the reaction was blocked at the p-quinone imine 20; and with 5-methyl-2-aminophenol (21) as substrate, the reaction was blocked at the dihydro-2-aminophenoxazinone 22. These findings suggested a mechanism in which 2-aminophenoxazinone formation proceeded via a quinone imine intermediate 4 that was trapped by a second molecule of 2-aminophenol. Oxidation of the adduct 5 to the p-quinone imine 6 was followed by a second conjugate addition and a final 2-electron oxidation to give the product, 2-aminophenoxazinone. The role of the enzyme in the catalysis of each of these steps was examined. It was found that the second conjugate addition generated a racemic center at C4a, suggesting that this reaction did not occur at the active site. A deuterium isotope effect on the cleavage of the C4-H bond of 2-aminophenol suggested that partial dissociation of an intermediate from the enzyme occurred after the first conjugate addition. It is proposed that 2-aminophenoxazinone synthesis proceeds via a sequence of three consecutive 2-electron aminophenol oxidations and that the aminophenol moiety is regenerated during the reaction sequence by facile tautomerization reactions. Thus, what initially appears to be an impressively complex mechanism may, in fact, be ingeniously simple.  相似文献   

5.
By reaction of aldolase with dihydroxyacetone phosphate an acid-labile intermediate is formed, which is in rapid equilibrium with the eneamine intermediate. The equilibrium concentration of the eneamine + the acid-labile intermediates is constant between pH 5.5 and 7.5 and is not significantly different for native and for carboxypeptidase-treated aldolase. These data are in keeping with the view that the CH bond breaking and the CH bond forming at the C3 of dihydroxyacetone phosphate are affected to the same extent by the carboxypeptidase treatment. The formation of the acid-labile intermediate is reversed by the addition of hexitol bisphosphate or by the removal of the dihydroxyacetone phosphate present in the medium; both these reactions display a biphasic time course. The acid-labile intermediate disappears rapidly when the enzyme-substrate complex is oxidized by ferricyanide, in this case the biphasic behavior is not observed. This means that practically all the acid-labile intermediate is rapidly converted into the eneamine and becomes available for the condensation reaction. At the equilibrium the enzyme-fructose bisphosphate and the enzyme-triose phosphate complexes represent 74 and 26%, respectively, of the total complexes, the rate constants for the condensation and for the cleavage reactions being, respectively, 19.3 and 6.7 s?1. These data support the view that the cleavage of the CC bond is the limiting step of the overall reaction.  相似文献   

6.
We investigated how the self-association of isolated tubulin dimers affects the rate of GTP hydrolysis and the equilibrium of nucleotide exchange. Both reactions are relevant for microtubule (MT) dynamics. We used HPLC to determine the concentrations of GDP and GTP and thereby the GTPase activity of SEC-eluted tubulin dimers in assembly buffer solution, free of glycerol and tubulin aggregates. When GTP hydrolysis was negligible, the nucleotide exchange mechanism was studied by determining the concentrations of tubulin-free and tubulin-bound GTP and GDP. We observed no GTP hydrolysis below the critical conditions for MT assembly (either below the critical tubulin concentration and/or at low temperature), despite the assembly of tubulin 1D curved oligomers and single-rings, showing that their assembly did not involve GTP hydrolysis. Under conditions enabling spontaneous slow MT assembly, a slow pseudo-first-order GTP hydrolysis kinetics was detected, limited by the rate of MT assembly. Cryo-TEM images showed that GTP-tubulin 1D oligomers were curved also at 36 °C. Nucleotide exchange depended on the total tubulin concentration and the molar ratio between tubulin-free GDP and GTP. We used a thermodynamic model of isodesmic tubulin self-association, terminated by the formation of tubulin single-rings to determine the molar fractions of dimers with exposed and buried nucleotide exchangeable sites (E-sites). Our analysis shows that the GDP to GTP exchange reaction equilibrium constant was an order-of-magnitude larger for tubulin dimers with exposed E-sites than for assembled dimers with buried E-sites. This conclusion may have implications on the dynamics at the tip of the MT plus end.  相似文献   

7.
Bromoperoxidase Compound I has been formed in reactions between bromoperoxidase and organic peroxide substrates. The absorbance spectrum of bromoperoxidase Compound I closely resembles the Compound I spectra of other peroxidases. The pH dependence of the second order rate constant for the formation of Compound I with hydrogen peroxide demonstrates the presence of an ionizable group at the enzyme active site having a pKa of 5.3. Protonation of this acidic group inhibits the rate of Compound I formation. This pKa value is higher than that determined for other peroxidases but the overall pH rate profiles for Compound I formation are similar. The one-electron reduction of bromoperoxidase Compound I yields Compound II and a second reduction yields native enzyme. Bromoperoxidase Compound II readily forms Compound III in the presence of an excess of hydrogen peroxide. Compound III passes through an as yet uncharacterized intermediate (III) in its decay to native enzyme. Compound III is produced and accumulates in enzymatic bromination reactions to become the predominate steady state form of the enzyme. Since Compound III is inactive as catalyst for enzymatic bromination, its accumulation leads to an idling reaction pathway which displays an unusual kinetic pattern for the bromination of monochlorodimedone.  相似文献   

8.
Human trypsin 4 is an unconventional serine protease that possesses an arginine at position 193 in place of the highly conserved glycine. Although this single amino acid substitution does not affect steady-state activity on small synthetic substrates, it has dramatic effects on zymogen activation, interaction with canonical inhibitors, and substrate specificity toward macromolecular substrates. To study the effect of a non-glycine residue at position 193 on the mechanism of the individual enzymatic reaction steps, we expressed wild type human trypsin 4 and its R193G mutant. 4-Methylumbelliferyl 4-guanidinobenzoate has been chosen as a substrate analogue, where deacylation is rate-limiting, and transient kinetic methods were used to monitor the reactions. This experimental system allows for the separation of the individual reaction steps during substrate hydrolysis and the determination of their rate constants dependably. We suggest a refined model for the reaction mechanism, in which acylation is preceded by the reversible formation of the first tetrahedral intermediate. Furthermore, the thermodynamics of these steps were also investigated. The formation of the first tetrahedral intermediate is highly exothermic and accompanied by a large entropy decrease for the wild type enzyme, whereas the signs of the enthalpy and entropy changes are opposite and smaller for the R193G mutant. This difference in the energetic profiles indicates much more extended structural and/or dynamic rearrangements in the equilibrium step of the first tetrahedral intermediate formation in wild type human trypsin 4 than in the R193G mutant enzyme, which may contribute to the biological function of this protease.  相似文献   

9.
CKX (cytokinin dehydrogenase) is a flavoprotein that cleaves cytokinins to adenine and the corresponding side-chain aldehyde using a quinone-type electron acceptor. In the present study, reactions of maize (Zea mays) CKX with five different substrates (N6-isopentenyladenine, trans-zeatin, kinetin, p-topolin and N-methyl-isopentenyladenine) were studied. By using stopped-flow analysis of the reductive half-reaction, spectral intermediates were observed indicative of the transient formation of a binary enzyme-product complex between the cytokinin imine and the reduced enzyme. The reduction rate was high for isoprenoid cytokinins that showed formation of a charge-transfer complex of reduced enzyme with bound cytokinin imine. For the other cytokinins, flavin reduction was slow and no charge-transfer intermediates were observed. The binary complex of reduced enzyme and imine product intermediate decays relatively slowly to form an unbound product, cytokinin imine, which accumulates in the reaction mixture. The imine product only very slowly hydrolyses to adenine and an aldehyde derived from the cytokinin N6 side-chain. Mixing of the substrate-reduced enzyme with Cu2+/imidazole as an electron acceptor to monitor the oxidative half-reaction revealed a high rate of electron transfer for this type of electron acceptor when using N6-isopentenyladenine. The stability of the cytokinin imine products allowed their fragmentation analysis and structure assessment by Q-TOF (quadrupole-time-of-flight) MS/MS. Correlations of the kinetic data with the known crystal structure are discussed for reactions with different cytokinins.  相似文献   

10.
Direct evidence for an enzyme-bound intermediate in the EPSP synthase reaction pathway has been obtained by rapid chemical quench-flow studies. The transient-state kinetic analysis has led to the following complete scheme: (formula; see text) Values for all 12 rate constants were obtained. Substrate trapping experiments in the forward and reverse reactions established the kinetically preferred order of binding and release of substrates and products and showed that shikimate 3-phosphate (S3P) and 5-enolpyruvoylshikimate 3-phosphate (EPSP) dissociate at rates greater than turnover in each direction. Pre-steady-state bursts of product formation were observed in the reaction in each direction indicating a rate-limiting step following catalysis. Single turnover experiments with enzyme in excess over substrate demonstrated the formation of a transient intermediate in both the forward and reverse reactions. In these experiments, the enzymatic reaction was observed by employing a radiolabel in the enol moiety of either phosphoenol pyruvate (PEP) or EPSP. The separation and quantitation of reaction products were accomplished by HPLC monitoring radioactivity. The intermediate was observed as the transient production of radiolabeled pyruvate, formed due to the breakdown of the intermediate in the acid quench used to stop the reaction. The intermediate was observed within 5-10 ms after the substrates were mixed with enzyme and decayed in a reaction paralleling the formation of product in each direction. Thus, the kinetics demonstrate directly the kinetic competence of the presumed intermediate. No pyruvate was formed, on a time scale which is relevant to catalysis, after incubation of the enzyme with dideoxy-S3P and PEP or with EPSP in the absence of phosphate; and so, the intermediate does not accumulate under these conditions. The intermediate broke down to form PEP and EPSP in addition to pyruvate when the reaction was quenched with base rather than acid; therefore, the intermediate must contain the elements of each product. Other experiments were designed to measure directly the phosphate binding rate and further constrain the PEP binding rate. The overall solution equilibrium constant in the forward direction was determined to be 180 by quantitation of radiolabeled reactants and products in equilibrium after incubation with a low enzyme concentration. The internal, active site equilibrium constant was obtained by incubation of radiolabeled S3P with excess enzyme and high concentrations of phosphate and PEP to provide the ratio of [EPSP]/[S3P] = 2.3, which is largely a measure of K4.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
The oxidative half-reactions of anthranilate hydroxylase (EC 1.14.12.2) were examined in the presence of anthranilate and modified substrates. C(4a)-Hydroperoxyflavin (C(4a)-FlOOH) and C(4a)-hydroxyflavin (C(4a)-FlOH) intermediates were detected in oxidative reactions with all substrates. Thus, the oxygenation reactions of the enzyme are similar to those of flavoprotein hydroxylases that convert phenolic compounds to catechols. These observations support a mechanism proposed for this enzyme (Powlowski, J. B., Dagley, S., Massey, V., and Ballou, D. P. (1987) J. Biol. Chem. 262, 69-74) involving nucleophilic attack of the substrate on C(4a)-FlOOH, and formation of an imine intermediate that is subsequently hydrolyzed. Anthranilate hydroxylase is therefore a typical flavoprotein hydroxylase with the added capacity of hydrolyzing imine intermediates. Fluorine substituents on the aromatic ring decreased the rate of conversion of C(4a)-FlOOH to C(4a)-FlOH, as predicted by this mechanism. Hydroxylation of 3-fluoro- and 3-methylanthranilates resulted in the formation of nonaromatic products that appeared to stabilize the C(4a)-FlOH. No evidence was found for a high extinction intermediate (intermediate II) (Entsch, B., Ballou, D. P., and Massey, V. (1976) J. Biol. Chem. 251, 2550-2563) under conditions where it was readily detected with other flavoprotein hydroxylases. It was shown that the spectra of the nonaromatic products (which are quinonoid forms) could not be summed with the spectra of C(4a)-hydroxyflavin to obtain that of a putative intermediate II, thus ruling out that explanation for previous observations of II.  相似文献   

12.
A convenient synthesis of novel bi/tricyclic azasugars fused thiazolidin-4-one and thiazinan-4-one by the one-pot tandem Staudinger/aza-Wittig/cyclization reaction under microwave radiation was demonstrated. The reactions were carried out with the azidosugar 1 and mercaptan acids via a key intermediate Schiff base and stereoselectively afforded the titled bi/tricyclic azasugars in good yield. All the dominant products were in the 1,2-trans form and the reaction stereoselectivity mainly depended upon the steric hindrance of the neighboring rigid cyclic isopropylidene groups on C-2, 3 which favors the exo-attack of the sulfur atom (in mercaptan acids) to the intermediate imine. The preliminary biological evaluation of the compounds 10-17 showed that compounds 10b, 11a, 12b, 14b, 16b, 17a, and 17b were found to active the natural killer (NK) cells significantly (immunopotentiating activity) and compounds 10a, 10b, 12a, 16b, and 17b exhibited weak inhibitory activity against β-glucosidase. Yet none of these tested compounds have obvious effects on T cell proliferation, or show inhibition against α-amylase and α-glucosidase.  相似文献   

13.
Nitrosobenzene (NOB) formed acid labile conjugates with reduced glutathione (GSH) and hemoglobin within red cells. In vitro, NOB rapidly reacted with GSH with formation of phenylhydroxylamine (PH), oxidized glutathione (GSSG), and a water-soluble compound identified as glutathionesulfinanilide (GSO-AN). Free aniline (AN), aminophenols and azoxybenzene were not detected. The proportion of PH formed increased with increasing GSH concentration and at higher pH values. Spectroscopic analysis revealed the formation of a labile adduct following a second order reaction (K = 5 x 10(3) M-1 . sec-1 at pH 7.4 and 37 degrees). This reaction was reversible because nearly all NOB could be extracted with ether from the labile intermediate. On the other hand, the labile intermediate was transformed into GSO-AN (with increasing rate at lower pH values) or it was cleaved by GSH with formation of GSSG and PH. Intermediate formation of NOB and thiol radicals was ruled out by analysis of the equilibrium data. A tentative scheme is presented for the proposed reaction mechanism.  相似文献   

14.
A1 hnRNP protein, which rapidly renatures complementary strands of nucleic acids in vitro, affects both the equilibrium and kinetic properties of the reaction (single-stranded DNA in equilibrium with double-stranded DNA). A1 lowers the melting transition of duplex DNA. However, at temperatures above this new Tm, both single- and double-stranded DNAs are present at equilibrium and are rapidly interconverting. Although the ratio of single and double strands under these conditions is a function of both the A1 protein and complementary DNA strand concentrations, it is not strongly affected by further increases in temperature. These surprising results demonstrate that A1 does not act as a simple catalyst in promoting renaturation and indicate how A1 and other proteins could act to speed the turnover of intermediate complexes in important biological processes.  相似文献   

15.
Microcalorimetry and high performance liquid chromatography have been used to conduct a thermodynamic investigation of reactions catalyzed by anthranilate synthase, the enzyme located at the first step in the biosynthetic pathway leading from chorismate to tryptophan. One of the overall biochemical reactions catalyzed by anthranilate synthase is: chorismate(aq) + ammonia(aq) = anthranilate(aq) + pyruvate(aq) + H2O(l). This reaction can be divided into two partial reactions involving the intermediate 2-amino-4-deoxyisochorismate (ADIC): chorismate(aq) + ammonia(aq) = ADIC(aq) + H2O(l) and ADIC(aq) = anthranilate(aq) + pyruvate(aq). The native anthranilate synthase and a mutant form of it that is deficient in ADIC lyase activity but has ADIC synthase activity were used to study the overall ammonia-dependent reaction and the first of the above two partial reactions, respectively. Microcalorimetric measurements were performed on the overall reaction at a temperature of 298.15 K and pH 7.79. Equilibrium measurements were performed on the first partial (ADIC synthase) reaction at temperatures ranging from 288.15 to 302.65 K, and at pH values from 7.76 to 8.08. The results of the equilibrium and calorimetric measurements were analyzed in terms of a chemical equilibrium model that accounts for the multiplicity of ionic states of the reactants and products. These calculations gave thermodynamic quantities at the temperature 298.15 K and an ionic strength of zero for chemical reference reactions involving specific ionic forms. For the reaction: chorismate2-(aq) + NH4+(aq) = anthranilate-(aq) + pyruvate-(aq) + H+(aq) + H2O(l), delta rHmo = -(116.3 +/- 5.4) kJ mol-1. For the reaction: chorismate2-(aq) + NH4+(aq) = ADIC-(aq) + H2O(l), K = (20.3 +/- 4.5) and delta rHmo = (7.5 +/- 0.6) kJ mol-1. Thermodynamic cycle calculations were used to calculate thermodynamic quantities for three additional reactions that are pertinent to this branch point of the chorismate pathway. The quantities obtained in this study permit the calculation of the position of equilibrium of these reactions as a function of temperature, pH, and ionic strength. Values of the apparent equilibrium constants and the standard transformed Gibbs energy changes delta rG'mo under approximately physiological conditions are given.  相似文献   

16.
Two colchicine analogs with modifications only in the C ring are better inhibitors than colchicine of cell growth and tubulin polymerization. Radiolabeled thiocolchicine (with a thiomethyl instead of a methoxy group at position C-10) and N-acetylcolchinol O-methyl ether (NCME) (with a methoxy-substituted benzenoid instead of the methoxy-substituted tropone C ring) were prepared for comparison with colchicine. Scatchard analysis indicated a single binding site with KD values of 1.0-2.3 microM. Thiocolchicine was bound 2-4 times as rapidly as colchicine, but the activation energies of the reactions were nearly identical (18 kcal/mol for colchicine, 20 kcal/mol for thiocolchicine). NCME bound to tubulin in a biphasic reaction. The faster phase was 60 times as fast as colchicine binding at 37 degrees C, and a substantial reaction occurred at 0 degrees C. The rate of the faster phase of NCME binding changed relatively little as a function of temperature, so the activation energy was only 7.0 kcal/mol. Dissociation reactions were also evaluated, and at 37 degrees C the half-lives of the tubulin-drug complexes were 11 min for NCME, 24 h for thiocolchicine, and 27 h for colchicine. Relative dissociation rates as a function of temperature varied little among the drug complexes. Activation energies for the dissociation reactions were 30 kcal/mol for thiocolchicine, 27 kcal/mol for NCME, and 24 kcal/mol for colchicine. Comparison of the activation energies of association and dissociation yielded free energies for the binding reactions of -20 kcal/mol for NCME, -10 kcal/mol for thiocolchicine, and -6 kcal/mol for colchicine. The greater effectiveness of NCME and thiocolchicine as compared with colchicine in biological assays probably derives from their more rapid binding to tubulin and the lower free energies of their binding reactions.  相似文献   

17.
M K Pangburn 《FEBS letters》1992,308(3):280-282
Purified alpha 2-macroglobulin and complement proteins C3 and C4 were treated with ammonia to break their intramolecular thioester bonds and reform the original free cysteinyl and glutamyl side chains. When this reaction was performed at low temperature a conformational intermediate was trapped which lacked a thioester, but which could refold to the native structure and spontaneously reform the thioester and full biological function. The findings suggest that these proteins may undergo spontaneous post-translational self-modification forming the thioesters without involvement of enzymes or high energy metabolites such as ATP.  相似文献   

18.
The kinetics of the reduction by aniline and a series of substituted anilines of a peroxidatically active intermediate, formed by oxidation of deuteroferriheme with hydrogen peroxide, have been studied by stopped-flow spectrophotometry. The reaction with aniline was first order with respect to [intermediate] and showed first-order saturation kinetics with respect to [aniline]. The second-order rate constant was 2.0 ± 0.2 × 105 M?1 sec?1 at 25°C (independent of pH in the range 6.60–9.68) compared with the value of 2.4 × 105 M?1 sec?1 for the reaction of aniline with horseradish peroxidase Compound I. The effect of aniline substituents upon reactivity towards the heme intermediate closely paralled those reported for reaction with the enzymic intermediate. Anilines bearing electron-donating substituents reacted more rapidly and those bearing electron-withdrawing substituents more slowly than the unsubstituted amine. The rate constants for the heme intermediate reactions (kdfh)found to be related to those for the enzymic reactions (khrp) by the equation:log kDFH= 0.65log kHRP+ 1.96 with a correlation coefficient of 0. 98.  相似文献   

19.
The aim of this study was to understand the mechanism of action through which carnosine (beta-alanyl-L-histidine) acts as a quencher of cytotoxic alpha,beta-unsaturated aldehydes, using 4-hydroxy-trans-2,3-nonenal (HNE) as a model aldehyde. In phosphate buffer solution (pH 7.4), carnosine was 10 times more active as an HNE quencher than L-histidine and N-acetyl-carnosine while beta-alanine was totally inactive; this indicates that the two constitutive amino acids act synergistically when incorporated as a dipeptide and that the beta-alanyl residue catalyzes the addition reaction of the histidine moiety to HNE. Two reaction products of carnosine were identified, in a pH-dependent equilibrium: (a) the Michael adduct, stabilized as a 5-member cyclic hemi-acetal and (b) an imine macrocyclic derivative. The adduction chemistry of carnosine to HNE thus appears to start with the formation of a reversible alpha,beta-unsaturated imine, followed by ring closure through an intra-molecular Michael addition. The biological role of carnosine as a quencher of alpha,beta-unsaturated aldehydes was verified by detecting carnosine-HNE reaction adducts in oxidized rat skeletal muscle homogenate.  相似文献   

20.
The 20 cysteine residues of tubulin are heterogeneously distributed throughout its three-dimensional structure. In the present work, we have used the reactivity of these cysteine residues with 5, 5'-dithiobis(2-nitrobenzoic acid) (DTNB) as a probe to detect the global conformational changes of tubulin under different experimental conditions. The 20 sulfhydryl groups can be classified into two categories: fast and slow reacting. Colchicine binding causes a dramatic decrease in the reactivity of the cysteine residues and causes complete protection of 1.4 cysteine residues. Similarly, other colchicine analogs that bind reversibly initially decrease the rate of reaction; but unlike colchicine they do not cause complete protection of any sulfhydryl groups. Interestingly, in all cases we find that all the slow reacting sulfhydryl groups are affected to the same extent, that is, have a single rate constant. Glycerol has a major inhibitory effect on all these slow reacting sulfhydryls, suggesting that the reaction of slow reacting cysteines takes place from an open state at equilibrium with the native. Ageing of tubulin at 37 degrees C leads to loss of self-assembly and colchicine binding activity. Using DTNB kinetics, we have shown that ageing leads to complete protection of some of the sulfhydryl groups and increased reaction rate for other slow reacting sulfhydryl groups. Ageing at 37 degrees C also causes aggregation of tubulin as indicated by HPLC analysis. The protection of some sulfhydryl groups may be a consequence of aggregation, whereas the increased rate of reaction of other slow reacting sulfhydryls may be a result of changes in global dynamics. CD spectra and acrylamide quenching support such a notion. Binding of 8-anilino-1-naphthalenesulfonate (ANS) and bis-ANS by tubulin cause complete protection of some cysteine residues as indicated by the DTNB reaction, but has little effect on the other slow reacting cysteines, suggesting local effects.  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