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.
The lysine-183 residues of yeast glyceraldehyde 3-phosphate dehydrogenase, in contrast to the cysteine-149 residues, react independently with acylating and alkylating agents. Modification of all four residues is required to inactivate the enzyme in spite of the fact that this residue is apparently in the neighborhood of the cysteine-149 involved in half-of-the-sites activity. The modification of the lysine-183 residue, however, influences the half-of-the-sites effect since alkylation of the cysteine-149 residues of the enzyme whose lysine-183 residues are acetylated follows a linear pattern with each subunit acting independently. Four lysine residues outside the active site can be modified with fluorodinitrobenzene, causing 80% loss in enzyme activity. Once again each subunit acts independently. This same residue can also be modified by a fluorescein label which can serve as a reporter group for binding and conformational changes occurring at the active site. The results add support for the functional symmetry of the apo-enzyme and demonstrate how the co-operativity between subunits can be altered by amino acid modification.  相似文献   

2.
A site-directed mutation, F235C, was created at the penultimate residue of the lambda-repressor. Measurement of dimer-monomer dissociation constant suggested that dimer-monomer dissociation of the mutant repressor is similar to that of the wild-type. Affinity towards a single operator O(R)1 is also similar to that of the wild-type repressor. The mutant repressor gene in a multi-copy plasmid confers immunity towards infection by a cI(-) lambda phage, suggesting preservation of functional integrity. Far-UV circular dichroism spectra show no major change in the secondary structure. Fluorescence quenching experiments, however, suggest increased exposure of some tryptophan residues. The urea denaturation profile indicates decreased stability of a part of the C-terminal domain. Under non-denaturing conditions, cysteine-235 shows half-of-the-sites reactivity, i.e. on average only one out of two cysteine-235 residues in the dimer shows reactivity towards sulfhydryl reagents. Fluorescence energy transfer between randomly labeled donor and acceptor fluorescent probes indicates that only one sulfhydryl per dimer is reactive, suggesting true half-of-the-sites reactivity. The structural role of the C-terminal tail in the whole repressor dimer is discussed.  相似文献   

3.
Pyrene maleimide is shown to be a 'half of the sites' reagent for glutamate dehydrogenase and for glyceraldehyde-3-phosphate dehydrogenase. The modified residues are identified as cysteine-115 for glutamate dehydrogenase and cysteine-149 for glyceraldehyde-3-phosphate dehydrogenase. The two enzymes react differently with pyrene maleimide. Whereas the hydrophobic environment of cysteine-115 directs the modification of glutamate dehydrogenase, the high reactivity of cysteine-149 determines the specific modification of glyceraldehyde-3-phosphate dehydrogenase. Glutamate dehydrogenase activity is unaltered by the modification: glyceraldehyde-3-phosphate dehydrogenase activity in inhibited.  相似文献   

4.
Glyceraldehyde 3-phosphate dehydrogenase (D-glyceraldehyde-3-phoshate:nicotinamide adenine dinucleotide oxidoreductase (phosphorylating), EC 1.2.1.12) forms a complex with 3-pyridinealdehyde-NAD which survives precipitation with 7% perchloric acid. The molar ratio bound 3-pyridinealdehyde-NAD to the enzyme is 2.5 to 2.9. Lactate, malate, and alcohol dehydrogenases do not form acid-precipitable complexes with 3-pyridinealdehyde-NAD. 3-Pyridinealdehyde-deamino-NAD or glyceraldehyde 3-phosphate also forms an acid-stable complex with glyceraldehyde 3-phosphate dehydrogenase; however, NAD, 3-acetylpyridine-NAD, or thionicotinamide-NAD does not produce an acid-stable complex. Incubation of the glyceraldehyde 3-phosphate dehydrogenase with glyceraldehyde 3-phosphate, acetyl phosphate, iodoacetic acid, or iodosobenzoate inhibits the formation of the acid-stable complex with 3-pyridinealdehyde-NAD. Glyceraldehyde 3-phosphate or 3-pyridinealdehyde-NAD also prevents carboxymethylation of the active site cysteine-149 by[14-C]iodoacetic acid. These studies indicate that the aldehyde group of 3-pyridinealdehyde-NAD forms a thiohemiacetal linkage with cysteine-149 which is the substrate binding site for the dehydrogenase reaction. These findings may account for the fact that 3-pyridinealdehyde-NAD strongly inhibits the dehydrogenase and esterase activities of 3-pyridinealdehyde-NAD forms a thiohemiacetal linkage with cysteine-149 which is the substrate binding site for the dehydrogenase reaction. These findings may account for the fact that 3-pyridinealdehyde-NAD strongly inhibits the dehydrogenase and esterase activities of glyceraldehyde 3-phosphate dehydrogenase which require reduced cysteine-149. However, the analogue does not inhibit the acetyl phosphates activity of the enzyme for which the active site sulfhydryl residues must be oxidized.  相似文献   

5.
A study of the kinetics of a heterodimeric variant of glutathione transferase (GST) A1-1 has led to the conclusion that, although the wild-type enzyme displays all-of-the-sites reactivity in nucleophilic aromatic substitution reactions, it demonstrates half-of-the-sites reactivity in addition reactions. The heterodimer, designed to be essentially catalytically inactive in one subunit due to a single point mutation (D101K), and the two parental homodimers were analyzed with seven different substrates, exemplifying three types of reactions catalyzed by glutathione transferases (nucleophilic aromatic substitution, addition, and double-bond isomerization reactions). Stopped-flow kinetic results suggested that the wild-type GST A1-1 behaved with half-of-the-sites reactivity in a nucleophilic aromatic substitution reaction, but steady-state kinetic analyses of the GST A1-D101K heterodimer revealed that this was presumably due to changes to the extinction coefficient of the enzyme-bound product. In contrast, steady-state kinetic analysis of the heterodimer with three different substrates of addition reactions provided evidence that the wild-type enzyme displayed half-of-the-sites reactivity in association with these reactions. The half-of-the-sites reactivity was shown not to be dependent on substrate size, the level of saturation of the enzyme with glutathione, or relative catalytic rate.  相似文献   

6.
The reaction of iodine with glyceraldehyde 3-phosphate dehydrogenase from Bacillus stearothermophilus was investigated. The active-site thiol group of the cysteine residue homologous with cysteine-149 in the pig muscle enzyme was protected by reaction with tetrathionate. The apoenzyme was readily inhibited by KI3 solution at pH8, but the coenzyme, NAD+, protected the enzyme against inhibition and decreased the extent of iodination. At pH 9.5, ready inhibition of both apo- and holo-enzyme was observed. Tryptic peptides containing residues iodinated at pH 8 were isolated and characterized. One of the most reactive residues in both holo- and apo-enzymes was a tyrosine homologous with tyrosine-46 in the pig muscle enzyme, and this residue was iodinated without loss of enzymic activity. Other reactive tyrosine residues in the apoenzyme were in positions homologous with residues 178, 273, 283 and 311 in the pig muscle enzyme, but they were not readily iodinated in the holoenzyme. Histidine residues in both holo- and apo-enzymes were iodinated at pH 8 in sequence positions homologous with residues 50, 162 and 190 in the pig muscle enzyme. The inhibition of the enzyme was not correlated with the iodination of a particular residue. The results are discussed in relation to a three-dimensional model based on the structure of the lobster muscle enzyme and demonstrate that conformational changes affecting the reactivity of several tyrosine residues most probably occur on binding of the coenzyme.  相似文献   

7.
Homotetrameric MSDH (methylmalonate semialdehyde dehydrogenase) from Bacillus subtilis catalyses the NAD-dependent oxidation of MMSA (methylmalonate semialdehyde) and MSA (malonate semialdehyde) into PPCoA (propionyl-CoA) and acetyl-CoA respectively via a two-step mechanism. In the present study, a detailed mechanistic characterization of the MSDH-catalysed reaction has been carried out. The results suggest that NAD binding elicits a structural imprinting of the apoenzyme, which explains the marked lag-phase observed in the activity assay. The enzyme also exhibits a half-of-the-sites reactivity, with two subunits being active per tetramer. This result correlates well with the presence of two populations of catalytic Cys302 in both the apo- and holo-enzymes. Binding of NAD causes a decrease in reactivity of the two Cys302 residues belonging to the two active subunits and a pKapp shift from approx. 8.8 to 8.0. A study of the rate of acylation as a function of pH revealed a decrease in the pKapp of the two active Cys302 residues to approx. 5.5. Taken to-gether, these results support a sequential Cys302 activation process with a pKapp shift from approx. 8.8 in the apo-form to 8.0 in the binary complex and finally to approx. 5.5 in the ternary complex. The rate-limiting step is associated with the b-decarboxylation process which occurs on the thioacylenzyme intermediate after NADH release and before transthioesterification. These data also indicate that bicarbonate, the formation of which is enzyme-catalysed, is the end-product of the reaction.  相似文献   

8.
Specific reaction of Cys-149 with 3,3,3-trifluorobromacetone allows one to probe symmetry relation between the active center regions of tetrameric glyceraldehyde-3-phosphate dehydrogenase by 19F nuclear magnetic resonance (nmr) techniques. Nmr titration studies in the pH range of greatest enzymic activity reveal the existence of species with an (alphaalpha')2 structure; this symmetry is not induced by the coenzyme. Addition of NADH to the ketone-labeled protein causes the enzymic reduction of the ligand in a stereospecific manner and is used to demonstrate the functionality of residues other than Cys-149 that are essential for catalysis. The interpretation of chemical shift characteristics found for the trifluoroacetonyl group together with the kinetics of its reduction allows the derivation of a dynamic model for the enzymic structure which may contribute to understanding of the half-of-the-sites phenomenon.  相似文献   

9.
Previous photolabeling and limited proteolysis studies suggested that one of the four basic residues (Arg-141) of the N-terminal cytoplasmic loop connecting helices IV and V (loop 4-5) of the melibiose permease (MelB) from Escherichia coli has a potential role in its symport function (Ambroise, Y., Leblanc, G., and Rousseau, B. (2000) Biochemistry 39, 1338-1345). A mutagenesis study of Arg-141 and of the other three basic residues of loop 4-5 was undertaken to further examine this hypothesis. Cys replacement analysis indicated that Arg-141 and Arg-149, but not Lys-138 and Arg-139, are essential for MelB transport activity. Replacement of Arg-141 by neutral residues (Cys or Gln) inactivated transport and energy-independent carrier-mediated flows of substrates (counterflow, efflux), whereas it had a limited effect on co-substrate binding. R141C sugar transport was partially rescued on reintroducing a positive charge with a charged and permeant thiol reagent. Whereas R149C was completely inactive, R149K and R149Q remained functional. Strikingly, introduction of an additional mutation in the C-terminal helix X (Gly for Val-343) of R149C restored sugar transport. Impermeant thiol reagents inhibited R149C/V343G transport activity in right-side-out membrane vesicles and prevented sugar binding in a sugar-protected manner. All these data suggest that MelB loop 4-5 is close to the sugar binding site and that the charged residue Arg-141 is involved in the reaction of co-substrate translocation or substrate release in the inner compartment.  相似文献   

10.
The aminoacylation reaction catalyzed by the dimeric tryptophanyl-tRNA synthetase from beef pancreas was studied under pre-steady-state conditions by the quenched-flow method. The transfer of tryptophan to tRNATrp was monitored by using preformed enzyme-bis(tryptophanyl adenylate) complex. Combinations of either unlabeled or L-[14C]tryptophan-labeled tryptophanyl adenylate and of aminoacylation incubation mixtures containing either unlabeled tryptophan or L-[14C]tryptophan were used. We measured either the formation of a single labeled aminoacyl-tRNATrp per enzyme subunit or the turnover of labeled aminoacyl-tRNATrp synthesis. Four models were proposed to analyze the experimental data: (A) two independent and nonequivalent subunits; (B) a single active subunit (subunits presenting absolute "half-of-the-sites reactivity"); (C) alternate functioning of the subunits (flip-flop mechanism); (D) random functioning of the subunits with half-of-the-sites reactivity. The equations corresponding to the formation of labeled tryptophanyl-tRNATrp under each labeling condition were derived for each model. By use of least-squares criteria, the experimental curves were fitted with the four models, and it was possible to disregard models B and C as likely mechanisms. Complementary experiments, in which there was no significant excess of ATP-Mg over the enzyme-adenylate complex, emphasized an activator effect of free L-tryptophan on the rate of aminoacylation. This result disfavored model A. Model D was in agreement with all data. The analyses showed that the transfer step was not the major limiting reaction in the overall aminoacylation process.  相似文献   

11.
12.
The sequence of peptides necessary to inhibit melittin-induced lysis was studied using 13 peptide analogues of the inhibitor Ac-IVIFDC-NH2. Although this inhibitor is a disulfide-linked dimer, inhibition was equally effective if the thiol SH was blocked or replaced by methionine or lysine. The substitution of phenylalanine with other aromatic residues preserved activity, as did the replacement of aspartic acid by asparagine. The results suggest that the cytolytic activity of melittin can be inhibited by a short peptide of four hydrophobic residues followed by two other nonspecific residues. Fluorescence studies showed that the inhibitor caused a blue shift in the Trp emission spectrum. A spin label attached to the N-terminus of the inhibitor significantly quenched the fluorescence. These data confirmed the involvement of Trp 19 with the inhibitor, also predicted by molecular modeling of the probable binding site. Density gradient studies with large unilamellar vesicles indicated that the inhibitor prevented melittin from reacting with the lipid bilayer.  相似文献   

13.
The complex between tetrameric glucose 6-phosphate dehydrogenase (G6PD) and four moles of structural NADP can bind four additional NADP equivalents with a Kdiss of 0.85 μM. Alternatively, this complex shows a maximal binding capacity of two NADPH equivalents, with a corresponding Kdiss of 0.1 μM. Therefore, a clear discrepancy has emerged from these spectrofluorimetric titrations with either the oxidized or the reduced form of the coenzyme, an “all-of-the-sites reactivity” being observed for NADP and a “half-of-the-sites reactivity” being conversely involved in NADPH binding.  相似文献   

14.
Wei Y  Lin J  Frey PA 《Biochemistry》2001,40(37):11279-11287
UDP-galactose 4-epimerase contains the coenzyme NAD+ bound tightly at the active site. NAD+ functions as the coenzyme for the interconversion of UDP-galactose and UDP-glucose by reversibly mediating their dehydrogenation to the common intermediate UDP-4-ketohexopyranoside. The epimerase structure and spectrophotometric data indicate that NAD+ may engage in electrostatic interactions with amino acid side chains that may regulate the reactivity of NAD+. In this work, we carried out NMR studies of [nicotinamide-4-13C]NAD+ bound to wild-type epimerase and epimerases mutated at amino acid residues in contact with NAD+. The 4-13C NMR chemical shifts revealed the following: The 4-13C chemical shift in wild-type epimerase is 149.9 ppm; mutation of Ser 124 to Ala changes it slightly by 0.2 ppm to 150.1 ppm; mutation of Tyr 149 to Phe results in a downfield perturbation of 2.7 ppm to 152.6 ppm; and the simultaneous mutation of Ser 124 to Ala and Tyr 149 to Phe also causes a downfield perturbation of 2.8 ppm to 152.7 ppm. Mutation of Lys 153 to Met results in a 13C chemical shift of 150.8 ppm, which is 0.9 ppm downfield from that of wild type and 1.8 ppm upfield from that of Y149F-epimerase. The 13C chemical shifts of nicotinamide C4 of NAD+ in these epimerases are correlated with their respective reactivities with NaBH3CN. In addition, reactivity of NAD+ in wild-type and S124A-epimerases displays pH dependence, with higher rates at lower pH where Tyr 149 in these two enzymes is protonated. The results support an electrostatic model in which repulsion between positively charged Lys 153 and N1 of the nicotinamide ring increases the reactivity of NAD+, while the phenolate of Tyr 149 opposes the positive electrostatic field and attenuates the reactivity of NAD+. Ser 124 has very little effect on the electron distribution within the nicotinamide ring or the reactivity of NAD+. The effects of binding the substrate analogue P1-uridyl-P2-methyl diphosphate (Me-UDP) on the 4-13C chemical shifts are opposite to those induced by the mutations. MeUDP perturbs the 4-13C chemical shift 2.9 ppm downfield in the wild-type and S124A-epimerases but has little or no effect in the cases of Y149F- or K153M-epimerases. The results support the postulate that NAD+ activation induced by uridine nucleotides is brought about by a conformational change of epimerase that repositions Tyr 149 at an increased distance from nicotinamide N1 of NAD+ while maintaining the electrostatic repulsion between Lys 153 and nicotinamide N1 of NAD+.  相似文献   

15.
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   

16.
Z Chen  L Lu  M Shirley  W R Lee  S H Chang 《Biochemistry》1990,29(5):1112-1118
Three amino acid residues (glycine-14, cysteine-135, and cysteine-218) previously speculated to be important for the structure and function of Drosophila melanogaster alcohol dehydrogenase have been investigated by using site-directed mutagenesis followed by kinetic analysis and chemical modification. Mutating glycine-14 to valine (G14V) virtually inactivates Drosophila ADH, and substitution of alanine at this position (G14A) causes a 31% decrease in activity. Thermal denaturation and kinetic and inhibition studies further demonstrate that replacing glycine-14 with either alanine or valine leads to structural changes in the NAD binding domain. These results provide direct evidence for the role played by glycine-14 in maintaining the correct conformation in the NAD binding domain. On the other hand, changing of cysteine-135, -218, or both to alanine (C135A, C218A, and C135A/C218A) causes no decrease in the catalytic activity of the enzyme, indicating that neither of the cysteinyl residues is essential for catalysis. C135A and wild-type enzyme are both inactivated by DTNB. In contrast, C218A and C135A/C218A are unaffected by DTNB treatment. DTNB modification of cysteine-218 can be prevented by the substrates NAD and 2-propanol, suggesting that cysteine-218 may be in the vicinity of the active site. Cysteine-135 which is normally insensitive to DTNB becomes accessible in the presence of 2-propanol and/or NAD, suggesting a conformational change induced by binding of these substrates.  相似文献   

17.
The primary structure of human C1 inhibitor was determined by peptide and DNA sequencing. The single-chain polypeptide moiety of the intact inhibitor is 478 residues (52,869 Da), accounting for only 51% of the apparent molecular mass of the circulating protein (104,000 Da). The positions of six glucosamine-based and five galactosamine-based oligosaccharides were determined. Another nine threonine residues are probably also glycosylated. Most of the carbohydrate prosthetic groups (probably 17) are located at the amino-terminal end (residues 1-120) of the protein and are particularly concentrated in a region where the tetrapeptide sequence Glx-Pro-Thr-Thr, and variants thereof, is repeated 7 times. No phosphate was detected in C1 inhibitor. Two disulfide bridges connect cysteine-101 to cysteine-406 and cysteine-108 to cysteine-183. Comparison of the amino acid and cDNA sequences indicates that secretion is mediated by a 22-residue signal peptide and that further proteolytic processing does not occur. C1 inhibitor is a member of the large serine protease inhibitor (serpin) gene family. The homology concerns residues 120 through the C-terminus. The sequence was compared with those of nine other serpins, and conserved and nonconserved regions correlated with elements in the tertiary structure of alpha 1-antitrypsin. The C1 inhibitor gene maps to chromosome 11, p11.2-q13. C1 inhibitor genes of patients from four hereditary angioneurotic edema kindreds do not have obvious deletions or rearrangements in the C1 inhibitor locus. A HgiAI DNA polymorphism, identified following the observation of sequence variants, will be useful as a linkage marker in studies of mutant C1 inhibitor genes.  相似文献   

18.
P-Glycoprotein (Pgp) is an important transport enzyme composed of two homologous domains and transports a wide range of structurally diverse xenobiotics from the cell. Recent studies have indicated that allosteric interactions occur between the nucleotide binding domains and between the substrate binding domains of the two halves, but the extent of this interaction as well as the means by which the enzyme can transport such a wide variety of substrates has not been elucidated. Herein, the Pgp-mediated transport of a marker substrate, daunorubicin (DNR), out of viable cells was examined in the presence of a variety of other known substrates of Pgp. For most of the typical Pgp substrates examined, the relationship between inhibition of DNR efflux and competing substrate concentration was sigmoidal and therefore not a simple mutually exclusive competitive inhibition of transport. The Hill coefficient ranged from about 3 to 5 for the inhibition of transport of DNR. This negative cooperativity in combination with recent evidence, including several examples of noncompetitive inhibition between the homologous halves of Pgp, indicates a "half-of-the-sites" reactivity. Our data support the mechanistic proposal that substrate binding at one putative transport binding site precludes activity at another unequal site; many of the substrates examined exert a negative allosteric effect on the other transport site (and vice versa). A half-of-the-sites reactivity model would account for many of these observations and may be critical to the efficiency of Pgp substrate transport of a broad spectrum of compounds.  相似文献   

19.
Analysis of the quantitative precipitation of bovine β-lactoglobulin (β-Lg) with rabbit antiserum to β-Lg indicated that there were at least four antigenic sites on β-Lg. To study the antigenic property of bovine β-Lg, we examined the antigenic reactivity of anti β-Lg serum with β-Lg specifically modified with chemical reagents by immunodiffusion analysis, a quantitative precipitin test, and an enzyme-linked immunosorbent assay.

Modification of arginine residues, tryptophan residues, or sulfhydryl groups had little effect on the antigenic reactivity. A significant decrease in the reactivity was observed when β-Lg was acetylated, succinylated, modified with diethylpyrocarbonate or coupled with glycine amide.

These results suggest that 1.1 of three arginine residues, two tryptophan residues, and one sulfhydryl group are out of the antigenic sites, but there is a possibility that the amino group, histidine residue and carboxyl group may play an important role in the antigenicity of bovine β-Lg.  相似文献   

20.
Site-specific mutagenesis has been used to prepare two mutant forms of the alpha subunit of tryptophan synthase from Salmonella typhimurium in which either cysteine-81 or cysteine-118 is replaced by a serine residue. These mutant proteins are potentially useful for x-ray crystallographic studies since a heavy metal binding site is specifically eliminated in each mutant. The purified mutant proteins are fully active in four reactions catalyzed by the wild type alpha 2 beta 2 complex of tryptophan synthase. However, the mutant alpha 2 beta 2 complexes dissociate more readily and are less heat-stable than the wild type alpha 2 beta 2 complex. Thus, cysteine-81 and cysteine-118 of the alpha subunit serve structural but not functional roles.  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