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.
meso-α,?-Diaminopimelate D-dehydrogenase was inhibited by sulfhydryl reagents such as p-chloromercuribenzoate and HgCl2. Two sulfhydryl groups were titrated per molecule in the presence and absence of 6 M guanidine hydrochloride: the enzyme contained one sulfhydryl group per subunit. Modification of the sulfhydryl groups with p-chloromercuribenzoate, 5,5'-dithiobis(2-nitrobenzoic acid), 4,4'-dithiopyridine, N-ethylmaleimide, and iodoacetic acid was accompanied by a loss of enzyme activity. However, modification of sulfhydryl groups of the enzyme with cyanide did not affect the activity. Thus, the introduction of bulky or charged substituents to sulfhydryl groups decreased the catalytic activity of the enzyme, but modification of the groups with the small and uncharged group, a cyano group, did not. The sulfhydryl groups did not play an essential role in catalysis.  相似文献   

2.
Formaldehyde dehydrogenase from Pseudomonas putida C-83 was found to contain 7 halfcystine residues per subunit monomer, as checked by the method of performic acid oxidation. Approximately 7 sulfhydryl groups per subunit monomer were titrated with 5,5′-dithiobis(2-nitrobenzoic acid) (DTNB) after denaturation with 8 m urea. In the native enzyme, modification of three sulfhydryl groups per subunit with p-chloromercuribenzoate (PCMB) led to the complete loss of enzyme actiyities for both formaldehyde and n-butanol. Hydrogen-peroxide competitively inhibited the enzyme activity for formaldehyde, while it was only slightly inhibitory to the activity for n-butanol. Both formaldehyde and hydrogen-peroxide protected one sulfhydryl group per subunit monomer from modification with PCMB. Moreover, hydrogen-peroxide was hardly reactive to the enzyme which was preincubated with formaldehyde.

From these observations, we conclude that one of three PCMB-reactive sulfhydryl groups is essential for the binding of formaldehyde, and hydrogen-peroxide modifies this sulfhydryl group.  相似文献   

3.
The acetyl-CoA:acetoacetate CoA-transferase of Escherichia coli has the subunit structure α2β2 The enzyme contains six sulfhydryl groups, one per α chain and two per β chain, and no disulfides. The rates and extent of sulfhydryl group reactivity with 5,5′-dithiobis(2-nitrobenzoic acid) were compared in the free enzyme, the enzyme-CoA intermediate in the catalytic pathway, and a substrate analog-enzyme Michaelis complex. The analog used was acetylaminodesthio-CoA, a competitive inhibitor with respect to acetyl-CoA; the analog is not a substrate. The reactions were studied in the presence and absence of 10% glycerol. In the absence of glycerol, one sulfhydryl group reacted rapidly in the free enzyme and enzyme-CoA intermediate; relative to the free enzyme, the rate and number of subsequently reacting sulfhydryl groups were increased in the enzyme-CoA intermediate. In the presence of 10% glycerol, one sulfhydryl group reacted rapidly in the free enzyme, while two reacted rapidly in the enzyme-CoA compound; the rates and extents of subsequently reacting sulfhydryl groups were also enhanced in the enzyme-CoA compound. The data strongly suggested subunit interactions in the free enzyme and intermediate; glycerol abolished those interactions in the enzyme-CoA intermediate. In the absence of glycerol, sulfhydryl group reactivity in the Michaelis complex, enzyme-acetylaminodesthio-CoA, was similar to that in the free enzyme with one exception: One of the more slowly reacting sulfhydryl groups in the free enzyme reacted at a rate characteristic of the enzyme-CoA intermediate. The results obtained with N-ethylmaleimide were qualitatively similar. The fractional inactivation of the enzyme with N-ethylmaleimide as a function of sulfhydryl groups modified and the subunit location of those sulfhydryl groups indicated that the same sulfhydryl groups react in both enzyme species; however, those sulfhydryl groups reacted more rapidly in the enzyme-CoA compound. The data indicate both subunit interactions in the enzyme and characteristic conformational changes upon formation of an acyl-CoA-enzyme Michaelis complex and the enzyme-CoA intermediate.  相似文献   

4.
The apoenzyme of diol dehydrase was inactivated by four sulfhydryl-modifying reagents, p-chloromercuribenzoate, 5,5′-dithiobis(2-nitrobenzoate) (DTNB), iodoacetamide, and N-ethylmaleimide. In each case pseudo-first-order kinetics was observed. p-Chloromercuribenzoate modified two sulfhydryl groups per enzyme molecule and modification of the first one resulted in complete inactivation of the enzyme. DTNB also modified two sulfhydryl groups, but modification of the second one essentially corresponded to the inactivation. In both cases, the inactivation was reversed by incubation with dithiothreitol. Cyanocobalamin, a potent competitive inhibitor of adenosylcobalamin, protected the essential residue, but not the nonessential one, against the modification by these reagents. By resolving the sulfhydryl-modified cyanocobalamin-enzyme complex, the enzyme activity was recovered, irrespective of treatment with dithiothreitol. From these results, we can conclude that diol dehydrase has two reactive sulfhydryl groups, one of which is essential for catalytic activity and located at or in close proximity to the coenzyme binding site. The other is nonessential for activity. Neitherp-chloromercuribenzoate- nor DTNB-modified apoenzyme was able to bind cyanocobalamin, whereas the iodoacetamide- and N-ethylmaleimide-modified apoenzyme only partially lost the ability to bind cyanocobalamin. The inactivation of diol dehydrase by p-chloromercuribenzoate and DTNB did not bring about dissociation of the enzyme into subunits. Total number of the sulfhydryl groups of this enzyme was 14 when determined in the presence of 6 m guanidine hydrochloride. No disulfide bond was detected.  相似文献   

5.
We previously found a very large NAD-dependent glutamate dehydrogenase with approximately 170?kDa subunit from Janthinobacterium lividum (Jl-GDH) and predicted that GDH reaction occurred in the central domain of the subunit. To gain further insights into the role of the central domain, several single point mutations were introduced. The enzyme activity was completely lost in all single mutants of R784A, K810A, K820A, D885A, and S1142A. Because, in sequence alignment analysis, these residues corresponded to the residues responsible for glutamate binding in well-known small GDH with approximately 50?kDa subunit, very large GDH and well-known small GDH may share the same catalytic mechanism. In addition, we demonstrated that C1141, one of the three cysteine residues in the central domain, was responsible for the inhibition of enzyme activity by HgCl2, and HgCl2 functioned as an activating compound for a C1141T mutant. At low concentrations, moreover, HgCl2 was found to function as an activating compound for a wild-type Jl-GDH. This suggests that the mechanism for the activation is entirely different from that for the inhibition.  相似文献   

6.
The aspergillic acid degrading enzyme (ADE) that catalyzes the cleavage of the pyrazine ring in aspergillic acid (AA, l-hydroxy-3-isobutyl-6-sec-butyl-2-pyrazinone) was purified to electrophoretic homogeneity from extracts of Trichoderma koningii ATCC 76666. ADE was a homodimeric protein with a molecular mass of 112kDa, contained lmol of FAD per mol of subunit, and required NAD(P)H and molecular oxygen for its activity. ADE had an isoelectric point of around 5.3, and an optimum pH of 7.0–8.0. p-Chloromercuribenzoate and HgCl2 completely inhibited ADE activity, while metal chelating reagents, α, α′-dipyridyl and o-phenanthroline, were not inhibitors. The substrate specificity among AA-related compounds was that hydroxyaspergillic acid was a poor substrate (16% of the activity for AA) and deoxyaspergillic acid did not serve as a substrate.  相似文献   

7.
Three mercuric chloride binding sites are identified on l-glutamate dehydrogenase. In the presence of EDTA, the binding of two mercuric chloride molecules per subunit induces the dissociation of the polyhexamers into hexamers. The physical and catalytic properties of this modified hexamer are similar to those of the native enzyme. This induced dissociation of the enzyme is probably the result of an exclusive binding of the mercurial to the free hexamer, and the dissociation velocity does not appear to be rate limited by the binding reaction of the mercurials. The third mercuric chloride binding site is protected by both EDTA and l-glutamate. The binding of HgCl2 to this site leads to the complete inactivation of the protein. There is no overlap between these modifications of l-glutamate dehydrogenase and the two previously described modifications of the enzyme by mercurial.  相似文献   

8.
Our findings that the apo β2 subunit of tryptophan synthase of Escherichia coli is inactivated by the modification of one sulfhydryl residue per monomer by nitrothiocyanobenzoic acid and is reactivated by removal of the CN group indicate that the reactive sulfhydryl residue (SH-I) is essential for catalytic activity. SH-I is shown to be the same residue which was previously found to react with bromoacetylpyridoxamine phosphate and different from a sulfhydryl (SH-II) which reacts with N-ethylmaleimide in the presence of pyridoxal phosphate. The results of partial tryptic digestions of β2 subunit labeled selectively at SH-I or SH-II show that both sulfhydryl residues are located in the F1 fragment which also contains the pyridoxal phosphate binding site.  相似文献   

9.
NAD+-dependent and NADP+-dependent glyceraldehyde-3-phosphate (G-3-P) dehydrogenases were isolated from Euglena gracilis and characterized as to their physical and chemical parameters. NAD+-G-3-P dehydrogenase was found to have a strong resemblance to similar enzymes from muscle tissue. It has a molecular weight of about 140,000, four subunits of identical size and charge, and a single species of NH2-terminal amino acid. Two sulfhydryl groups per subunit are present, one of which is directly involved in the catalytic activity and is rapidly titratable. The enzyme also exhibits the “half the sites reactivity” of sulfhydryl groups as defined by O. P. Malhotra and S. A. Bernhard ((1968) J. Biol. Chem. 243, 1243). The pH and temperature optima are also similar to those of the enzymes from muscle tissue, as are the reaction kinetics and the strict specificity for NAD+.NADP+-dependent G-3-P dehydrogenase is different in many respects. Its molecular weight is slightly lower (~136,000) than that of the NAD+ enzyme, though it also consists of four subunits. It has a higher affinity for the reverse reaction substrates, in line with its probable function in vivo in CO2 fixation. There is only one sulfhydryl group per subunit, and that is not involved in activity, suggesting a difference in reaction mechanisms between the two enzymes. The NADP+-dependent enzyme exhibits activation by ATP, whereas the NAD+-dependent enzyme is competitively inhibited by this nucleotide.The greatest difference observed is in the physical characteristics of the enzymes. NADP+-G-3-P dehydrogenase was highly hydrophobic. Its solubility in a 10% aqueous solution of p-dioxane was approximately four to five times that of the NAD+-enzyme. Isolation of the enzyme was accomplished by fractionation in 1,2-dimethoxyethane, which also stabilized the enzymatic activity, as did aqueous p-dioxane. The high axial ratio of the NADP+-enzyme (~9) coupled with its very low degree of hydration as well as the high degree of amidation of the dicarboxylic amino acids (>90%) indicates that the exterior of the enzyme molecule is probably hydrophobic in nature. This is in agreement with its in vivo hydrophobic environment in the chloroplast membrane and explains the lability of the enzyme once extracted into an aqueous environment as well as its stabilization in solvents.  相似文献   

10.
The stoichiometry and the reactivity of the sulfhydryl groups of a human liver acid phosphatase have been studied. The smallest (Mr = 14,400) of the three molecular-weight forms of acid phosphatase from human liver, recently purified and characterized in our laboratory, was treated with various sulfhydryl group-specific reagents: p-hydroxymercuribenzoate, p-hydroxymercuriphenylsulfonate, fluorescein mercuriacetate, methyl methanethiosulfonate, p-nitrophenoxycarbonyl methyl disulfide, and thiosulfate. A total loss of enzymatic activity was obtained in each case. By spectrophotometric titration with 5,5′-dithiobis(2-nitrobenzoate) and p-hydroxymercuriphenylsulfonate it was shown that there are six free sulfhydryls per protein molecule, consistent with the amino acid analysis of this enzyme. The same number was deduced as a result of inactivation studies carried out with p-hydroxymercuribenzoate and p-hydroxymercuriphenylsulfonate. A total loss of activity was obtained at reagent to enzyme ratios of 6:1 in both cases. Similar results were obtained upon inactivation by p-nitrophenoxycarbonyl methyl disulfide, where the enzyme was found to possess only 10% residual activity at an inhibitor-to-enzyme ratio of 6:1. With fluorescein mercuriacetate as an inactivator, total loss of activity was found at a 2.5 times molar excess of this reagent over protein. Both the stoichiometry of inactivation and fluorescence titration experiments suggest that fluorescein mercuriacetate can function as a bifunctional sulfhydryl group reagent. The activity of a totally inactivated enzyme preparation obtained following reaction with excess of p-nitrophenoxycarbonyl methyl disulfide or with methyl methanethiolsulfonate could be almost completely restored upon treatment with dithiothreitol. These data are consistent with the interpretation that in each enzyme molecule, there are six free sulfhydryl groups of almost equal reactivity, at least one of which is essential for enzymatic activity.  相似文献   

11.
The structure of soybean β-amylase in trigonal (P3221) crystals was determined at 4.5 Å resolution by X-ray crystallographic techniques using the isomorphous replacement method. X-Ray diffraction data were collected by the screened precession method for the native enzyme and two heavy atom derivatives. The shape of the enzyme molecule and the locations of mercurial binding are presented. The molecule appeared to be composed of two domains: the larger domain contains one mercurial site on its surface and the smaller domain has another mercurial site, which seemed to be the so-called essential sulfhydryl group. A distinct cleft formed between the domains near the latter sulfhydryl group may be a substrate binding region.  相似文献   

12.
Diamine oxidase from Euphorbia characias latex contains two sulfhydryl groups per mole of dimeric enzyme. The sulfhydryl groups are unreactive in the native enzyme but can be readily titrated by 4,4′-dithiodipyridine after protein denaturation, or anaerobically in the presence of the amine substrate. In the presence of both substrates (diamine and oxygen) they react sluggishly. The sulfhydryl groups show different reactivity toward various reagents, but in every case their modification inhibits catalytic activity. The insensitivity of the native enzyme to specific reagents suggests that the sulfhydryl groups are positioned in the interior of the protein and shielded from the solvent. Their reactivity in the presence of the amine substrate could be attributed to a conformational change occurring upon substrate binding or after substrate oxidation.  相似文献   

13.
Tetrahedral anions bind to a phosphate binding site of 6-phosphogluconate dehydrogenase from Candida utilis, inhibit the enzyme competitively with the 6-phosphogluconate, decrease the reactivity of the SH groups, and mimic the protective effect of 6-phosphogluconate against some inactivating agents. The reaction of the enzyme with butanedione results in the inactivation of the enzyme associated with the modification of a single arginine residue per subunit. This arginine residue may be involved in the binding of the phosphate to the enzyme. Inactivation of the enzyme, upon reaction with permanganate, appears to be due to the oxidation to cysteic acid of a single cysteine residue per enzyme subunit. The reaction of the enzyme with either periodate or hexachloroplatinate causes the loss of the catalytic activity. This inactivation, due to an affinity labeling, is correlated with the oxidation of two SH groups per subunit to an S-S bridge. Photoinactivation of the enzyme by pyridoxal 5′-phosphate is also restricted to the active site of the enzyme. The lysine and the histidine residues involved in this photoinactivation should thus be in the vicinity of the phosphate binding site.  相似文献   

14.
In the presence of Hg2+Ascaris lipoamide dehydrogenase stimulated the reduction of oxygen, ferricyanide, and 2,6-dichlorophenolindophenol with NADH, which was inhibited by lipoic acid. On the other hand, Cu2+ stimulated the reduction of the artificial dyes, but only a little the reduction of oxygen. Hg2+ changed the visible absorption spectrum of the lipoamide dehydrogenase, but did not change the fluorescence curve. Lipoic acid decreased the fluorescence, but did not change the visible absorption spectrum. The Ascaris lipoamide dehydrogenase have two SH groups per one subunit and 5–6 moles of HgCl2 and 3–4 moles of CuSO4 per one subunit were required for the maximal activity.  相似文献   

15.
The free catalytic subunit of cAMP-dependent protein kinase readily undergoes a pronounced, salt-induced conformational change at neutral pH and around physiological values of ionic strength. This change, which is fully reversible, can be monitored directly by the relative chemical reactivity of two SH groups in the enzyme. Upon increasing the ionic strength of the medium from 0.03 to 0.22, one sulfhydryl becomes more reactive towards 5,5′-dithiobis[2-nitrobenzoic acid] while the other sulfhydryl becomes less reactive towards the same reagent. In parallel, the enzyme undergoes a salt-induced inactivation when histone H2b is used as a substrate. Though not reflected in the Vmax, this conformational change considerably increases the Km of the enzyme for histone H2b as well as for MgATP. This intrinsic malleability of the enzyme can account for the well-known salt inhibition of the enzyme for certain substrates and ion-dependent activation towards other substrates. It is suggested that this malleability might constitute the molecular basis for modulating the specificity of the enzyme and channeling its activity from one substrate to another in response to intracellular specifier signals.  相似文献   

16.
《Plant science》1986,45(1):1-7
NADP-malate dehydrogenase (L-malate: NADP+ oxidoreductase, E.C. 1.1.1.82) was purified from the leaves of Zea mays L. and its subunit molecular weight, amino acid composition and the changes in number of thiol groups during activation were determined. The amino acid composition we found differed from that reported earlier for the Z. mays enzyme but was very similar to that reported for the enzyme isolated from pea leaves. The maize enzyme contains fewer methionine residues (3 compared to 5 in pea) but a greater total number of cysteine residues (6 compared to 3 in pea). In its inactive form (oxidised) the enzyme contained 2 thiols per subunit of which only 1 reacts with 5,5′-dithiobis(2-nitrobenzoic acid) when the enzyme is in its native form. During activation by dithiothreitol two disulphide bonds are reduced per subunit to give 4 new thiol groups. We conclude that NADP-malate dehydrogenase from leaves of the C4 plant Z. mays is very similar to the enzyme from the C3 plant pea. However, apparently two disulphide bonds are reduced during the reductive activation of the Z. mays enzyme in vitro compared with 1 disulphide bond for the pea enzyme.  相似文献   

17.
Liver microsomal cytochrome P-448 purified from 3-methylcholanthrene-treated rats or rabbits contained seven free sulfhydryl groups per mole of enzyme as determined by amino acid analysis or by spectrophotometric titrations with 5,5′-dithiobis(2-nitroben-zoic acid), 4,4′-dipyridinedisulfide, 2-nitro-5-thiocyanobenzoic acid, and p-mercuribenzoate. The rat cytochrome P-448-catalyzed hydroxylation of benzo[a]pyrene was inhibited 70% after modification of the enzyme with 5,5′-dithiobis(2-nitrobenzoic acid) but was unaffected after titration of the enzyme with other sulfhydryl reagents, suggesting that the sulfhydryl groups may not be essential for catalysis. On the other hand, the rabbit cytochrome P-448-catalyzed hydroxylation of benzo[a]pyrene was inhibited following the modification of this enzyme with all of the sulfhydryl reagents listed above. Whether the loss in catalytic activity in this case is due to the essential role of the sulfhydryl groups in catalysis or to the steric hindrance or conformational change due to the substituent is uncertain.  相似文献   

18.
The chemical modification of cysteine and tyrosine residues in formyltetrahydrofolate synthetase from Clostridium thermoaceticum has been examined relative to enzymatic activity and reactivity of these groups in the native protein. 4,4′-Dipyridyl disulfide, dansylaziridine, and fluorescein mercuric acetate all reacted with just one of six sulfhydryls per enzyme subunit, resulting in activities of 100, 95 and 70%, respectively. The Km values for MgATP, formate, and tetrahydrofolate were unaltered in the modified enzymes. ATP did produce a 2.5-fold reduction in the rate of reaction between the enzyme and 4,4′-dipyridyl disulfide. Tetranitromethane reacted most rapidly with a single sulfhydryl group per subunit to produce a 20–30% loss in activity. Subsequent additions of tetranitromethane modified 2.2 tyrosines per subunit which was proportional to the loss of the remaining enzymatic activity. Folic acid, a competitive inhibitor, protected against modification of the tyrosines and the associated activity losses; however, the oxidation of the single sulfhydryl group and the initial 20–30% activity loss were unaffected. In the presence of folic acid, higher concentrations of tetranitromethane produced a loss of the remaining activity proportional to the modification of 1.2 tyrosines per subunit. It is proposed that at least 1 tyrosine critical for enzymatic activity is located at or near the folic acid/tetrahydrofolate binding site.  相似文献   

19.
Aspartase (L-aspartate ammonia-lyase, EC 4.3.1.1) of Escherichia coli W contains 38 half-cystine residues per tetrameric enzyme molecule. Two sulfhydryl groups were modified with N-ethylmaleimide or 5,5'-dithiobis(2-nitrobenzoic acid) (DTNB) per subunit, while 8.3 sulfhydryl groups were titrated with p-mercuribenzoic acid. In the presence of 4 M guanidine - HCl, 8.6 sulfhydryl groups reacted with DTNB per subunit. Aspartase was inactivated by various sulfhydryl reagents following pseudo-first-order kinetics. Upon modification of one sulfhydryl group per subunit with N-Ethylmaleimide, 85% of the original activity was lost; a complete inactivation was attained concomitant with the modification of two sulfhydryl groups. These results indicate that one or two sulfhydryl groups are essential for enzyme activity. L-Aspartate and DL-erythro-beta-hydroxyaspartate markedly protected the enzyme against N-ethylmaleimide-inactivation. Only the compounds having an amino group at the alpha-position exhibited protection, indicating that the amino group of the substrate contributes to the protection of sulfhydryl groups of the enzyme. Examination of enzymatic properties after N-ethylmaleimide modification revealed that 5-fold increase in the Km value for L-aspartate and a shift of the optimum pH for the activity towards acidic pH were brought about by the modification, while neither dissociation into subunits nor aggregation occurred. These results indicate that the influence of the sulfhydryl group modification is restricted to the active site or its vicinity of the enzyme.  相似文献   

20.
Gustine DL 《Plant physiology》1981,68(6):1323-1326
White clover (Trifolium repens L.) callus tissue cultures accumulated the phytoalexin medicarpin after treatment with sulfhydryl reagents. After 24-hour exposures to sulfhydryl reagents, maximum obtainable levels of medicarpin, determined by high performance liquid chromatography analysis, were found with 50 millimolar N-ethyl maleimide, 25 millimolar HgCl2, 2 millimolar p-chloromercuribenzoic acid, and 0.5 millimolar iodoacetamide. Increased medicarpin levels were also observed in callus treated with p-chloromercuribenzene sulfonic acid, but the highest concentration tested (11.8 millimolar) did not produce the maximum response. After sulfhydryl treatment, medicarpin levels were unchanged for 4 to 6 hours, but steadily increased thereafter with maximum accumulation occurring by 48 to 50 hours for p-chloromercuribenzoic acid, p-chloromercuribenzene sulfonic acid, and HgCl2 treated callus. Medicarpin levels did not increase in iodoacetamide-treated callus until 8 hours after sulfhydryl exposure, and medicarpin levels were still increasing linearly after 50 hours. Three other metabolic inhibitors, KCN, NaF, and Na3AsO4, did not exhibit elicitor activity, indicating cell death was not a factor in the response. Pretreatment of callus with 20 millimolar dithiothreitol followed by 40 millimolar N-ethyl maleimide did not produce the phytoalexin response. Preincubation with dithiothreitol also prevented elicitor activity of HgCl2 and p-chloromercuribenzene sulfonic acid. These results suggested that dithiothreitol pretreatment somehow prevented sulfhydryl groups within the cell from reacting with the test compounds. These experiments established that the integrity of sulfhydryl groups is important in regulating phytoalexin accumulation in callus cells.  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