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.
Allicin—diallyl thiosulfinate—is the main biologically active component of freshly crushed garlic. Allicin was synthesized as described elsewhere and was tested for its inhibitory ability against jack bean urease in 20?mM phosphate buffer, pH 7.0 at 22°C. The results indicate that allicin is an enzymatic inactivator. The loss of urease activity was irreversible, time- and concentration dependent and the kinetics of the inactivation was biphasic; each phase, obeyed pseudo-first-order kinetics. The rate constants for inactivation were measured for the fast and slow phases and for several concentrations of allicin. Thiol reagents, and competitive inhibitor (boric acid) protected the enzyme from loss of enzymatic activity. The studies demonstrate that urease inactivation results from the reaction between allicin and the SH-group, situated in the urease active site (Cys592).  相似文献   

2.
At low pH, EDTA promotes the loss of the tightly bound nickel ions from jack bean urease. The specific activity of soluble enzyme after partial EDTA-promoted inactivation is a linear function of the nickel content. The results are consistent with the presence of 2.0 nickel ions per 97 000-dalton subunit in pure urease. The time scale for loss of enzymatic activity and nickel under these conditions is similar to that for loss of the "abnormal" tail absorption in the ultraviolet and visible absorption spectrum of urease (including the shoulder at approximately 420 nm). This indicates that nickel in urease is essential for enzymatic activity and establishes that the metal ions are in part responsible for the tail absorption in the ultraviolet spectrum of urease. After partial inactivation in the presence of EDTA either at low pH or in 2.5 M guanidinium chloride at neutral pH, urease did not regain activity in the presence of Ni2+. As yet apourease has not been produced reversibly. Jack bean seeds grown hydroponically without added nickel were low in both urease activity and nickel (10 and 6%, respectively, of parent seeds). Several other metal ions were readily available. This result suggests that metal ions other than nickel cannot substitute for nickel in the formation of normally active urease.  相似文献   

3.
The kinetics of Klebsiella aerogenes urease inactivation by disulfide and alkylating agents was examined and found to follow pseudo-first-order kinetics. Reactivity of the essential thiol is affected by the presence of substrate and competitive inhibitors, consistent with a cysteine located proximal to the active site. In contrast to the results observed with other reagents, the rate of activity loss in the presence of 5,5'-dithiobis(2-nitrobenzoic acid) (DTNB) saturated at high reagent concentrations, indicating that DTNB must first bind to urease before inactivation can occur. The pH dependence for the rate of urease inactivation by both disulfide and alkylating agents was consistent with an interaction between the thiol and a second ionizing group. The resulting macroscopic pKa values for the 2 residues are less than 5 and 12. Spectrophotometric studies at pH 7.75 demonstrated that 2,2'-dithiodipyridine (DTDP) modified 8.5 +/- 0.2 mol of thiol/mol of enzyme or 4.2 mol of thiol/mol of catalytic unit. With the slow tight binding competitive inhibitor phenyl-phosphorodiamidate (PPD) bound to urease, 1.1 +/- 0.1 mol of thiol/mol of catalytic unit were protected from modification. PPD-bound DTDP-modified urease could be reactivated by dialysis, consistent with the presence of one thiol per active site. Analogous studies at pH 6.1, using the competitive inhibitor phosphate, confirmed the presence of one protected thiol per catalytic unit. Under denaturing conditions, 25.5 +/- 0.3 mol of thiol/mol of enzyme (Mr = 211, 800) were modified by DTDP.  相似文献   

4.
Reaction ofKlebsiella aerogenes urease with diethylpyrocarbonate (DEP) led to a pseudo-first-order loss of enzyme activity by a reaction that exhibited saturation kinetics. The rate of urease inactivation by DEP decreased in the presence of active site ligands (urea, phosphate, and boric acid), consistent with the essential reactive residue being located proximal to the catalytic center. ThepH dependence for the rate of inactivation indicated that the reactive residue possessed apK a of 6.5, identical to that of a group that must be deprotonated for catalysis. Full activity was restored when the inactivated enzyme was treated with hydroxylamine, compatible with histidinyl or tyrosinyl reactivity. Spectrophotometric studies were consistent with DEP derivatization of 12 mol of histidine/mol of native enzyme. In the presence of active site ligands, however, approximately 4 mol of histidine/mol of protein were protected from reaction. Each protein molecule is known to possess two catalytic units; hence, we propose that urease possesses at least one essential histidine per catalytic unit.  相似文献   

5.
Reaction ofKlebsiella aerogenes urease with diethylpyrocarbonate (DEP) led to a pseudo-first-order loss of enzyme activity by a reaction that exhibited saturation kinetics. The rate of urease inactivation by DEP decreased in the presence of active site ligands (urea, phosphate, and boric acid), consistent with the essential reactive residue being located proximal to the catalytic center. ThepH dependence for the rate of inactivation indicated that the reactive residue possessed apK a of 6.5, identical to that of a group that must be deprotonated for catalysis. Full activity was restored when the inactivated enzyme was treated with hydroxylamine, compatible with histidinyl or tyrosinyl reactivity. Spectrophotometric studies were consistent with DEP derivatization of 12 mol of histidine/mol of native enzyme. In the presence of active site ligands, however, approximately 4 mol of histidine/mol of protein were protected from reaction. Each protein molecule is known to possess two catalytic units; hence, we propose that urease possesses at least one essential histidine per catalytic unit.  相似文献   

6.
Pyrocatechol was studied as an inhibitor of jack bean urease in 20 mM phosphate buffer, pH 7.0, 25 degrees C. The inhibition was monitored by an incubation procedure in the absence of substrate and reaction progress studies in the presence of substrate. It was found that pyrocatechol acted as a time- and concentration dependent irreversible inactivator of urease. The dependence of the residual activity of urease on the incubation time showed that the rate of inhibition increased with time until there was total loss of enzyme activity. The inactivation process followed a non-pseudo-first order reaction. The obtained reaction progress curves were found to be time-dependent. The plots showed that the rate of the enzyme reaction in the final stages reached zero. From protection experiments it appeared that thiol-compounds such as L-cysteine, 2-mercaptoethanol and dithiothreitol prevented urease from pyrocatechol inactivation as well as the substrate, urea, and the competitive inhibitor boric acid. These results proved that the urease active site was involved in the pyrocatechol inactivation.  相似文献   

7.
Pyrocatechol was studied as an inhibitor of jack bean urease in 20?mM phosphate buffer, pH 7.0, 25°C. The inhibition was monitored by an incubation procedure in the absence of substrate and reaction progress studies in the presence of substrate. It was found that pyrocatechol acted as a time- and concentration dependent irreversible inactivator of urease. The dependence of the residual activity of urease on the incubation time showed that the rate of inhibition increased with time until there was total loss of enzyme activity. The inactivation process followed a non-pseudo-first order reaction. The obtained reaction progress curves were found to be time-dependent. The plots showed that the rate of the enzyme reaction in the final stages reached zero. From protection experiments it appeared that thiol-compounds such as l-cysteine, 2-mercaptoethanol and dithiothreitol prevented urease from pyrocatechol inactivation as well as the substrate, urea, and the competitive inhibitor boric acid. These results proved that the urease active site was involved in the pyrocatechol inactivation.  相似文献   

8.
N-Ethylmaleimide (NEM) was studied as an inactivator of jack bean urease at 25 degrees C in 20 mM phosphate buffer, pHs 6.4, 7.4, and 8.3. The inactivation was investigated by incubation procedure in the absence of a substrate. It was found that NEM acted as a time and concentration dependent inactivator of urease. The dependence of urease residual activity on the incubation time showed that the activity decreased with time until the total loss of enzyme activity. The process followed a pseudo-first-order reaction. A monophasic loss of enzyme activity was observed at pH 7.4 and 8.4, while a biphasic reaction occurred at pH 6.4. Moreover, the alkaline pH promoted the inactivation. The presence of thiol-compounds, such as L-cysteine, glutathione or dithiothreitol (DTT), in the incubation mixture significantly slowed down the rate of inactivation. The interaction test showed that the decrease of inactivation was an effect of NEM-thiol interaction that lowered NEM concentration in the incubation mixture. The reactivation of NEM-blocked urease by DTT application and multidilution did not result in an effective activity regain. The applied DTT reacted with the remaining inactivator and could stop the progress of enzyme activity loss but did not cause the reactivation. This confirmed the irreversibility of inactivation. Similar results obtained at pH 6.4, 7.4 and 8.4 indicated that the mechanism of urease inactivation by NEM was pH-independent. However, the pH value significantly influenced the process rate.  相似文献   

9.
N-Ethylmaleimide (NEM) was studied as an inactivator of jack bean urease at 25 °C in 20 mM phosphate buffer, pHs 6.4, 7.4, and 8.3. The inactivation was investigated by incubation procedure in the absence of a substrate. It was found that NEM acted as a time and concentration dependent inactivator of urease. The dependence of urease residual activity on the incubation time showed that the activity decreased with time until the total loss of enzyme activity. The process followed a pseudo-first-order reaction. A monophasic loss of enzyme activity was observed at pH 7.4 and 8.4, while a biphasic reaction occurred at pH 6.4. Moreover, the alkaline pH promoted the inactivation. The presence of thiol-compounds, such as L-cysteine, glutathione or dithiothreitol (DTT), in the incubation mixture significantly slowed down the rate of inactivation. The interaction test showed that the decrease of inactivation was an effect of NEM-thiol interaction that lowered NEM concentration in the incubation mixture. The reactivation of NEM-blocked urease by DTT application and multidilution did not result in an effective activity regain. The applied DTT reacted with the remaining inactivator and could stop the progress of enzyme activity loss but did not cause the reactivation. This confirmed the irreversibility of inactivation. Similar results obtained at pH 6.4, 7.4 and 8.4 indicated that the mechanism of urease inactivation by NEM was pH-independent. However, the pH value significantly influenced the process rate.  相似文献   

10.
During reaction with [14C]iodoacetamide at pH 6.3, radioactivity was incorporated primarily into a single Klebsiella aerogenes urease peptide concomitant with activity loss. This peptide was protected from modification at pH 6.3 by inclusion of phosphate, a competitive inhibitor of urease, which also protected the enzyme from inactivation. At pH 8.5, several peptides were alkylated; however, modification of one peptide, identical to that modified at pH 6.3, paralleled activity loss. The N-terminal amino acid sequence and composition of the peptide containing the essential thiol was determined. Previous enzyme inactivation studies of K. aerogenes urease could not distinguish whether one or two essential thiols were present per active site (Todd, M. J., and Hausinger, R. P. (1991) J. Biol. Chem. 266, 10260-10267); we conclude that there is a single essential thiol present and identify this residue as Cys319 in the large subunit of the heteropolymeric enzyme.  相似文献   

11.
Treatment of homogeneous preparations of Escherichia coli 2-keto-4-hydroxyglutarate aldolase with 1,2-cyclohexanedione, 2,3-butanedione, phenylglyoxal, or 2,4-pentanedione results in a time- and concentration-dependent loss of enzymatic activity; the kinetics of inactivation are pseudo-first order. Cyclohexanedione is the most effective modifier; a plot of log (1000/t 1/2) versus log [cyclohexanedione] gives a straight line with slope = 1.1, indicating that one molecule of modifier reacts with each active unit of enzyme. The kinetics of inactivation are first order with respect to cyclohexanedione, suggesting that the loss of activity is due to modification of 1 arginine residue/subunit. Controls establish that this inactivation is not due to modifier-induced dissociation or photoinduced structural alteration of the aldolase. The same Km but decreased Vmax values are obtained when partially inactivated enzyme is compared with native. Amino acid analyses of 95% inactivated aldolase show the loss of 1 arginine/subunit with no significant change in other amino acid residues. Considerable protection against inactivation is provided by the substrates 2-keto-4-hydroxyglutarate and pyruvate (75 and 50%, respectively) and to a lesser extent (40 and 35%, respectively) by analogs like 2-keto-4-hydroxybutyrate and 2-keto-3-deoxyarabonate. In contrast, formaldehyde or glycolaldehyde (analogs of glyoxylate) under similar conditions show no protective effect. These results indicate that an arginine residue is required for E. coli 2-keto-4-hydroxyglutarate aldolase activity; it most likely participates in the active site of the enzyme by interacting with the carboxylate anion of the pyruvate-forming moiety of 2-keto-4-hydroxyglutarate.  相似文献   

12.
Incubation of an NAD+-dependent succinic semialdehyde dehydrogenase from bovine brain with 4-dimethylaminoazobenzene-4-iodoacetamide (DABIA) resulted in a time-dependent loss of enzymatic activity. This inactivation followed pseudo first-order kinetics with a second-order rate constant of 168 m(-1).min(-1). The spectrum of DABIA-labeled enzyme showed a characteristic peak of the DABIA alkylated sulfhydryl group chromophore at 436 nm, which was absent from the spectrum of the native enzyme. A linear relationship was observed between DABIA binding and the loss of enzyme activity, which extrapolates to a stoichiometry of 8.0 mol DABIA derivatives per mol enzyme tetramer. This inactivation was prevented by preincubating the enzyme with substrate, succinic semialdehyde, but not by preincubating with coenzyme NAD+. After tryptic digestion of the enzyme modified with DABIA, two peptides absorbing at 436 nm were isolated by reverse-phase HPLC. The amino acid sequences of the DABIA-labeled peptides were VCSNQFLVQR and EVGEAICTDPLVSK, respectively. These sites are identical to the putative active site sequences of other brain succinic semialdehyde dehydrogenases. These results suggest that the catalytic function of succinic semialdehyde dehydrogenase is inhibited by the specific binding of DABIA to a cysteine residue at or near its active site.  相似文献   

13.
When ferric ion was added to solutions of the enzyme dextransucrase, first-order followed by second-order inactivation behavior was observed. The initial rapid activity loss was attributed to a ferric ion interacting with the thiol group of the native monomer to form a less active enzyme-ion complex; the second inactivation stage involved enzyme-ion complex aggregation and disulfide cross-link formation. In contrast, Cu2+ ion inactivation demonstrated simple first-order kinetics. As with Fe3+, Cu2+ ions can form complexes with enzyme thiol groups. However, unlike ferric ions, cupric ions can also strongly interact with the imidazole ring of histidine. Since the dextransucrase active site contains two key histidines, imidazole-cupric-ion interactions could potentially inhibit enzymatic activity. Thus, it was hypothesized that first-order Cu2+ inactivation kinetics involved the adsorption of this ion to the enzyme's activity site. The addition of a reducing agent such as dithiothreitol can inhibit the second enzyme aggregation stage by breaking disulfide cross-links but cannot restrict the initial formation of metal-enzyme complexes.  相似文献   

14.
δ-Aminolevulinic acid dehydratase (EC 4.2.1.24) was obtained in highly purified form from beef liver. Upon photooxidation of the enzyme in the presence of methylene blue as a sensitizer led to a loss of the enzymatic activity according to pseudo-first order kinetics. The pronounced pH dependence (pk value of 6.8) of the photooxidation rate and the results of amino acid analysis suggested that the inactivation was largely due to the modification of the histidine residue. The finding of the enzyme with little activity in the presence of diethylpyrocarbonate was consistent with such a speculation. On the basis of these results, it can be postulated that the histidine residue seems to play an important role in the enzymatic activity of δ-aminolevulinic acid dehydratase.  相似文献   

15.
Transketolase from baker's yeast is rapidly inactivated in the presence of 1-ethyl-3 (3'-dimethylaminopropyl)-carbodiimide or Woodward's reagent K. In both cases the kinetics of inactivation is biphasic, which agrees with the presence of two active centers in the enzyme molecule differing in their sensitivity to the inhibitors. There is some evidence that inactivation of transketolase is due to modification of carboxyl groups of enzyme. Complete inactivation is achieved by modification of one carboxyl per active site of the enzyme. The experimental results suggest that the carboxyl group is essential for the enzymatic activity of transketolase.  相似文献   

16.
Sarcosine oxidase [sarcosine: oxygen oxidoreductase (demethylating) EC 1.5.3.1] from Corynebacterium contained 8 sulfhydryl groups per mol of enzyme as determined with 5,5'-dithiobis(2-nitrobenzoic acid) (DTNB) in the presence of 0.2% SDS and by titration with p-chloromercuribenzoate (PMB). Among them, 2 groups were easily modified by iodoacetamide (IAA) and the modification resulted in complete loss of enzymatic activity. The inactivation by IAA followed first-order kinetics with respect to IAA concentration. The presence of acetate, a competitive inhibitor (I), protected the enzyme from inactivation by IAA. However, the protection was only approximately 50%. The enzyme was also inactivated by PMB, but in this case, there was practically no recovery of activity after treatment with thiol compounds. The enzyme was also rapidly inactivated by incubation with diethylpyrocarbonate (DEP). The absorbance change accompanying the inactivation showed that a single histidyl residue was modified by DEP, resulting in a complete loss of enzymatic activity. In the presence of acetate, the enzyme was completely protected from DEP-inactivation. Furthermore, DEP-inactivated enzyme recovered its enzymatic activity on treatment with hydroxylamine. These observations seem to imply that the modified histidine is essential for enzyme activity. In addition, modification by DEP changed the absorption spectrum in the visible region. This strongly suggests that the modified histidyl residue is present in the vicinity of the flavin moiety of the enzyme molecule.  相似文献   

17.
A model previously developed to characterize enzymatic in activation behavior was used to explain the non-first-order biphasic and grace period phenomena that are often observed with oligomeric enzymes. Luciferase and urease were used as model enzyme such as luciferase, the oligomer initially dissociates reversibly into two native monomer species. These native monomers can then reversibly denature and irreversibly aggregate and coagulate. With the hexamer, urease, two trimers are formed that can subsequently aggregate to form an inactive hexamer. The dissociated monomer species of luciferase do not possess catalytic activity, so the inactivation mechanism, is biphasic; the first slope of a first-order kinetic plot is influenced by the reversible oligomer/monomer/denatured-monomer transition. Whereas the second slope is associated with either irreversible aggregation or coagulation. In contrast, the trimer of urease has the same activity as the hexamer; therefore, during the intitial hexamer-trimer transition, little activity loss occurs. However, as the trimer concentration increases, activity decreases as a result of trimer aggregation. As a result, grace period inactivation behavior is observed. (c) 1992 John Wiley & Sons, Inc.  相似文献   

18.
The interaction of Escherichia coli glutamine synthetase with the adenosine 5'-triphosphate analogue, 5'-p-fluorosulfonylbenzoyladenosine (5'-FSO2BzAdo), has been studied. This interaction results in the covalent attachment of the 5'-FSO2BzAdo to the enzyme with concomitant loss of catalytic activity. Although adenine nucleotides interact with glutamine synthetase at three distinct sites--a noncovalent AMP effector site, a regulatory site of covalent adenylylation, and the catalytic ATP/ADP binding site--our studies suggest that reaction with 5'-FSO2BzAdo occurs only at the active center. When glutamine synthetase was incubated with 5'-FSO2BzAdo, the decrease in catalytic activity obeyed pseudo-first order kinetics. The plot of the observed rate constant of inactivation versus the concentration of 5'-FSO2BzAdo was hyperbolic, consistent with reversible binding of the analogue to the enzyme prior to covalent attachment. Protection against inactivation was afforded by ATP and ADP; L-glutamate did not protect the enzyme against inactivation, but rather enhanced the rate of inactivation, consistent with the observations of others (Timmons, R. B., Rhee, S. G., Luterman, D. L., and Chock, P. B. (1974) Biochemistry 13, 4479-4485) that there is synergism in the binding of the two substrates to the enzyme. The incorporation of approximately 1.09 mol of the 5'-FSO2BzAdo/mol of glutamine synthetase subunit resulted in the total loss of enzymatic activity. The results suggest that 5'-FSO2BzAdo occupies the ATP binding site at the active center of glutamine synthetase and binds covalently to an amino acid residue nearby.  相似文献   

19.
Alkaline phosphatase from Megalobatrachus japonicus was inactivated by diethyl pyrocarbonate (DEP). The inactivation followed pseudo-first-order kinetics with a second-order rate constant of 176 M(-1) x min(-1) at pH 6.2 and 25 degrees C. The loss of enzyme activity was accompanied with an increase in absorbance at 242 nm and the inactivated enzyme was re-activated by hydroxylamine, indicating the modification of histidine residues. This conclusion was also confirmed by the pH profiles of inactivation, which showed the involvement of a residue with pK(a) of 6.6. The presence of glycerol 3-phosphate, AMP and phosphate protected the enzyme against inactivation. The results revealed that the histidine residues modified by DEP were located at the active site. Spectrophotometric quantification of modified residues showed that modification of two histidine residues per active site led to complete inactivation, but kinetic stoichiometry indicated that one molecule of modifier reacted with one active site during inactivation, probably suggesting that two essential histidine residues per active site are necessary for complete activity whereas modification of a single histidine residue per active site is enough to result in inactivation.  相似文献   

20.
The reaction between N-ethylmaleimide and (Na+ + K+)-ATPase, performed under ligand conditions which produce each of the kinetic states of the enzyme and their associated conformational forms, was examined through an analysis of the inhibition of enzymatic activity and the incorporation of radiolabeled reagent into the enzyme. The inactivation reactions displayed pseudo-first order kinetics with respect to the concentration of active enzyme, indicating that the loss of activity is associated with the alkylation of a unique sulfhydryl group. In the absence of enzyme phosphorylation, the nucleophilicity of this sulfhydryl group is affected primarily by the nature of the monovalent cation present and does not correlate with the conformational state. A method for determining the actual concentration and specific radioactivity of radiolabeled N-ethylmaleimide during the reaction with (Na+ + K+)-ATPase was developed, allowing the measurement of the total reactive sulfhydryl groups of native (Na+ + K+)-ATPase under conditions identical with those of the inactivation studies. The labeling of the enzyme complex is associated almost exclusively with the large polypeptide, which contains four sulfhydryl groups which react with this reagent. One of these residues is presumably the sulfhydryl responsible for inactivation of the enzyme. Two react stoichiometrically and rapidly with N-ethylmaleimide under all conditions. The nucleophilicity of the fourth sulfhydryl group is governed by the conformational state of the enzyme, but the alkylation of this residue does not result in loss of enzymatic activity.  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