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 78 毫秒
1.
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   

2.
Fluorescein isothiocyanate (FITC) reactivity with the (Na+ + K+)-ATPase was studied at pH 6.5 and 9.0. Reaction with FITC is nearly complete in 30 min and is irreversible at both pH values. Differential inhibition of enzyme activity is observed at the two pH values as follows: at pH 6.5 the maximal inhibition reached is only 35–45% of the ATPase or p-nitrophenylphosphatase activities, whereas at pH 9.0 ATPase activity can be completely inhibited while maximal phosphatase inhibition is ca. 50%. At all concentrations of FITC tested, more FITC is incorporated into the enzyme at pH 9.0 than at 6.5. At both pH values NaCl increases the inhibition due to FITC while KCl protects against the inhibition. ATP protects the enzyme at both pH values with a K0.5 in the range of 8–20 μm. Enzyme that is partially inactivated at either pH shows no significant change in the K0.5 values for Na+ or K+ or in the Km app for ATP or p-nitrophenylphosphate for the remaining activity. The binding of 48VO4 is not changed by reaction with FITC at either pH, while [3H]ouabain binding is inhibited after reaction at pH 9.0 only in the presence of Mg+2 + Na+ + ATP. [3H]Ouabain binding in the presence of Mg+2 + inorganic phosphate is not inhibited by FITC reaction. Enzyme reacted at both pH values exhibits the expected fluorescein fluorescence (λex = 490, λem = 520) but only with enzyme reacted at pH 9.0 is fluorescence quenching by K+ or reversal by Na+ observed. These results suggest that different classes of amino groups react with FITC at the two pH values tested, and that these groups have distinct roles in the different activities of the enzyme.  相似文献   

3.
Light-induced interaction of Fe(II) cations with the donor side of Mn-depleted photosystem II (PS II(–Mn)) results in the binding of iron cations and blocking of the high-affinity (HAZ) Mn-binding site. The pH dependence of the blocking was measured using the diphenylcarbazide/2,6-dichlorophenolindophenol test. The curve of the pH dependence is bell-shaped with pK 1 = 5.8 and pK 2 = 8.0. The pH dependence of the O2-evolution mediated by PS II membranes is also bellshaped (pK 2 = 7.6). The pH dependence of the process of electron donation from exogenous donors in PS II(–Mn) was studied to determine the location of the alkaline pH sensitive site of the electron transport chain. The data of the study showed that the decrease in the iron cation binding efficiency at pH > 7.0 during blocking was determined by the donor side of the PS II(–Mn). Mössbauer spectroscopy revealed that incubation of PS II(–Mn) membranes in a buffer solution containing 57Fe(II) + 57Fe(III) was accompanied by binding only Fe(III) cations. The pH dependence of the nonspecific Fe(III) cation binding is also described by the same bell-shaped curve with pK 2 = 8.1. The treatment of the PS II(–Mn) membranes with the histidine modifier diethylpyrocarbonate resulted in an increase in the iron binding strength at alkaline pH. It is suggested that blocking efficiency at alkaline pH is determined by competition between OH and histidine ligand for Fe(III). Because the high-affinity Mn-binding site contains no histidine residue, this fact can be regarded as evidence that histidine is located at another (other than high-affinity) Fe(III) binding site. In other words, this means that the blockage of the high-affinity Mn-binding site is determined by at least two iron cations. We assume that inactivation of oxygen-evolving complex and inhibition of photoactivation in the alkaline pH region are also determined by competition between OH and a histidine residue involved in coordination of manganese cation outside the high-affinity site.  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.
ATPase activity of proton-translocating FOF1-ATP synthase (F-type ATPase or F-ATPase) is suppressed in the absence of protonmotive force by several regulatory mechanisms. The most conservative of these mechanisms found in all enzymes studied so far is allosteric inhibition of ATP hydrolysis by MgADP (ADP-inhibition). When MgADP is bound without phosphate in the catalytic site, the enzyme lapses into an inactive state with MgADP trapped.In chloroplasts and mitochondria, as well as in most bacteria, phosphate prevents MgADP inhibition. However, in Escherichia coli ATP synthase ADP-inhibition is relatively weak and phosphate does not prevent it but seems to enhance it.We found that a single amino acid residue in subunit β is responsible for these features of E. coli enzyme. Mutation βL249Q significantly enhanced ADP-inhibition in E. coli ATP synthase, increased the extent of ATP hydrolysis stimulation by sulfite, and rendered the ADP-inhibition sensitive to phosphate in the same manner as observed in FOF1 from mitochondria, chloroplasts, and most aerobic\photosynthetic bacteria.  相似文献   

6.
Treatment of an isoenzyme of potato apyrase of high adenosine triphosphatase/adenosine diphosphatase (ATPase/ADPase) ratio with iodine, N-acetylimidazole or tetranitromethane inactivates the ATPase activity of this enzyme faster than its ADPase activity. There was protection by substrates with the two last-named substances. This and the appearance of nitrotyrosine suggests the participation of tyrosyl residues in both enzymic activities of potato apyrase. The participation of thiol groups is excluded by the insensitivity of apyrase to p-chloromercuribenzoate. Also, 2-hydroxy-5-nitrobenzyl bromide or carboxymethylation produce the same rate of inactivation of ATPase and ADPase activities. Substrates protect both activities from inactivation. Hydrogen peroxide and photo-oxidation inactivate ATPase activity faster than ADPase activity. There is no protection by substrates. Analysis of pH effects on Vmax. and Km suggest different pK values for the amino acid residues at the ATP and ADP sites.  相似文献   

7.
The effect of the natural ATPase inhibitor and octylguanidine on the ATPase activity of soluble oligomycin-insensitive mitochondrial F1 were compared. Both compounds induced a maximal inhibition of 60–80% in various preparations of F1 studied. The inhibition was of the uncompetitive type with respect to MgATP, and the action of the compounds was partially additive. The data suggest that octylguanidine reproduces the action of the natural ATPase inhibitor. Alkylammonium salts also affect the ATPase activity in a similar form. F1 bound to Sepharose-hexylammonium is largely inactive, whilst free hexylammonium at higher concentrations induces only a partial inhibition of the activity. This suggests that the degree of immobilization of F1 is related to the magnitude of inhibition of ATPase activity induced by alkyl cations. The binding of F1 to Sepharose-hexylammonium is prevented by high concentrations of Na+ or K+.  相似文献   

8.
It was found that modification of thiol (SH-) groups of membrane proteins by Ellman’s reagent (5,5′-dithiol-bis-(2-nitrobenzoic) acid) results in inhibition of proton efflux and K+ influx in anaerobically grown (pH 7.5) wild-type strains of Escherichia coli and causes disturbances in K+-dependent, N,N′-dicyclohexylcarbodiimide-sensitive ATPase activity and molecular hydrogen production. No such effects were observed after substitution of the cysteine residue in the b-subunit of F0 of proton F0F1-ATPase for alanine. Moreover, the redox potential (RP) decreased as a result of H2 release during glucose fermentation and formate utilization was partly restored in the presence of Ellman’s reagent. Similar changes were established when another specific SH-reagent, succinimidyl-6(β-maleimidopropionamido)hexanoate, was used. Another thiol reagent, N-ethylmaleimide, did not exert such effects despite its inhibitory action on ion transport and ATPase activity. The data obtained provide conclusive evidence in favor of essential role of thiol groups and the cysteine residue in the b-subunit of F0 of F0F1-ATPase in proton-potassium exchange and H2 production in E. coli cells. The results also point to a possible involvement of SH-groups in the TrkA system of K+ uptake and an involvement of hydrogenases 3 or 4 in the interactions of these integral proteins with each other.  相似文献   

9.
Ribonuclease A has been trinitrophenylated to varying degrees by reaction with trinitrobenzenesulfonic acid. The reactive amino groups were identified by use of the peptides obtained from the oxidized TNP-RNase by tryptic and chymotryptic hydrolysis. From a quantitative study of the TNP-peptides it was possible to associate each amino group with values of pKma. It was shown that the lys-41 amino group had a pKa of 9.03 in TEA buffer. The pKa values of all of the other amino groups were dependent on the nature of the buffer (triethanolamine and phosphate) and on the pH.  相似文献   

10.
The pH profile of the rate of isomerization of 4,2′,4′-trihydroxychalcone catalyzed by chalcone isomerase shows dependence on the basic form of a group with a pK of 7.25. The same pH dependence is seen for the reverse reaction. Enzyme activity is lost in the presence of diethylpyrocarbonate at pH 6.0. In the presence of 20% formamide in imidazole buffers, the pK for the forward reaction is modified by a second pK of 7.1. This behavior represents a perturbed pK of a neutral acid group and is attributable to the 2′ hydroxyl of the chalcone substrate. These results suggest a mechanism of enzyme action involving nucleophilic addition of an imidazole group in the active site to the double bond followed by nucleophilic attack by the 2′ phenolate group, resulting in ring closure with inversion of configuration at C-2.  相似文献   

11.
Philip G. Koga  Richard L. Cross 《BBA》1982,679(2):269-278
1. Soluble beef-heart mitochondrial ATPase (F1) was incubated with [3H]pyridoxal 5′-phosphate and the Schiffbase complex formed was reduced with sodium borohydride. Spectral measurements indicate that lysine residues are modified and gel electrophoresis in the presence of detergent shows the tritium label to be associated with the two largest subunits, α and β. 2. In the absence of protecting ligands, the loss of ATP hydrolysis activity is linearly dependent on the level of pyridoxylation with complete inactivation correlating to 10 mol pyridoxamine phosphate incorporated per mol enzyme. Partial inactivation of F1 with pyridoxal phosphate has no effect on either the Km for ATP or the ability of bicarbonate to stimulate residual hydrolysis activity, suggesting a mixed population of fully active and fully inactive enzyme. 3. In the presence of excess magnesium, the addition of ADP or ATP, but not AMP, decreases the rate and extent of modification of F1 by pyridoxal phosphate. The non-hydrolyzable ATP analog, 5′-adenylyl-β, γ-imidodiphosphate, is particularly effective in protecting F1 against both modification and inactivation. Efrapeptin and Pi have no effect on the modification reaction. 4. Prior modification of F1 with pyridoxal phosphate decreases the number of exchangeable nucleotide binding sites by one. However, pyridoxylation of F1 is ineffective in displacing endogenous nucleotides bound at non-catalytic sites and does not affect the stoichiometry of Pi binding. 5. The ability of nucleotides to protect against modification and inactivation by pyridoxal phosphate and the loss of one exchangeable nucleotide site with the pyridoxylation of F1 suggest the presence of a positively charged lysine residue at the catalytic site of an enzyme that binds two negatively charged substrates.  相似文献   

12.
The K+-dependent p-nitrophenylphosphatase activity catalyzed by purified (Na+ + K+)-ATPase from pig kidney shows substrate inhibition (Ki about 9.5 mM at 2.1 mM Mg2+). Potassium antagonizes and sodium favours this inhibition. In addition, K+ reduces the apparent affinity for substrate activation, whereas p-nitrophenyl phosphate reduces the apparent affinity for K+ activation. In the absence of Mg2+, p-nitrophenyl phosphate, as well as ATP, accelerates the release of Rb+ from the Rb+ occluded unphosphorylated enzyme. With no Mg2+ and with 0.5 mM KCl, trypsin inactivation of (Na+ + K+)-ATPase as a function of time follows a single exponential but is transformed into a double exponential when 1 mM ATP or 5 mM p-nitrophenyl phosphate are also present. In the presence of 3 mM MgCl2, 5 mM p-nitrophenyl phosphate and without KCl the trypsin inactivation pattern is that described for the E1 enzyme form; the addition of 10 mM KCl changes the pattern which, after about 6 min delay, follows a single exponential. These results suggest that (i) the shifting of the enzyme toward the E1 state is the basis for substrate inhibition of the p-nitrophenulphosphatase acitivy of (Na+ + K+)-ATPase, and (ii) the substrate site during phosphatase activity is distinct from the low-affinity ATP site.  相似文献   

13.
H+-FOF1-ATP synthase (F-ATPase, F-type ATPase, FOF1 complex) catalyzes ATP synthesis from ADP and inorganic phosphate in eubacteria, mitochondria, chloroplasts, and some archaea. ATP synthesis is powered by the transmembrane proton transport driven by the proton motive force (PMF) generated by the respiratory or photosynthetic electron transport chains. When the PMF is decreased or absent, ATP synthase catalyzes the reverse reaction, working as an ATP-dependent proton pump. The ATPase activity of the enzyme is regulated by several mechanisms, of which the most conserved is the non-competitive inhibition by the MgADP complex (ADP-inhibition). When ADP binds to the catalytic site without phosphate, the enzyme may undergo conformational changes that lock bound ADP, resulting in enzyme inactivation. PMF can induce release of inhibitory ADP and reactivate ATP synthase; the threshold PMF value required for enzyme reactivation might exceed the PMF for ATP synthesis. Moreover, membrane energization increases the catalytic site affinity to phosphate, thereby reducing the probability of ADP binding without phosphate and preventing enzyme transition to the ADP-inhibited state. Besides phosphate, oxyanions (e.g., sulfite and bicarbonate), alcohols, lauryldimethylamine oxide, and a number of other detergents can weaken ADP-inhibition and increase ATPase activity of the enzyme. In this paper, we review the data on ADP-inhibition of ATP synthases from different organisms and discuss the in vivo role of this phenomenon and its relationship with other regulatory mechanisms, such as ATPase activity inhibition by subunit ε and nucleotide binding in the noncatalytic sites of the enzyme. It should be noted that in Escherichia coli enzyme, ADP-inhibition is relatively weak and rather enhanced than prevented by phosphate.  相似文献   

14.
Selective chemical modification was used to examine amino acid residues that might be critical for the operation of the gastric K+-stimulated ATPase. Modification of amino groups with the fluorigenic reagent 2-methoxy-2,4-diphenyl-3-dihydrofuranone resulted in selective inhibition of the K+-stimulated ATPase and H+-transporting activities of the gastric microsomes, while the Mg2+-ATPase was not affected. Half-maximal inhibition occurred at about 3 μg 2-methoxy-2,4-diphenyl-3-dihydrofuranone/ml at pH 8.5. ATP provided complete protection against inhibition; the apparent Km for ATP protection was about 50 μM. Nucleotide selectivity for protection was ATP > ADP > ITP > GTP > CTP > AMP. Sodium dodecyl sulfate gel electrophoresis of the reacted microsomes showed that virtually all the fluorescent label was on the Mr 100 000 peptide band, a very small peptide, and aminolipids. In the presence of ATP there was about 75% reduction in the fluorescent label on the Mr 100 000 peptide, but no change in the labeling of the other components. The arginine specific reagent, butanedione, inhibited Mg2+-ATPase and K+-ATPase activities, with the former being much less reactive. Similar to 2-methoxy-2,4-diphenyl-3-dihydrofuranone, ATP provided complete protection from butanedione treatment. It is concluded that amino and guanidino groups are critical to the function of the K+-ATPase and may be actually at the ATP binding site.  相似文献   

15.
The effect of phosphate buffer on the activity of jack bean urease was studied in the range of pH 5.80–8.07. The inhibition constants of phosphate buffer were determined by measuring initial reaction rates at each pH for a series of buffer concentrations at a series of urea concentrations. It was shown that: (1) at pH 5.80–7.49 the buffer is a competitive inhibitor of the enzyme with Ki,buffer increasing from 0.54 mM for pH 5.80 to 362 mM for pH 7.49, (2) the values of pKi,buffer are pH-dependent exhibiting a slope of −1 at pH 5.80–6.5 and a slope of −2 at pH 6.5–7.49, (3) from pH 7.62 as the pH is further raised the competitive inhibition of urease by the buffer was not observed, (4) the true competitive inhibitor of urease is H2PO4 ion, and (5) pH 6.5 and 7.6 correspond to the ionization constants of the active site groups of urease responsible for the inhibitory strength of H2PO4 ion.  相似文献   

16.
Phosphoenolpyruvate carboxylase from the Crassulacean acid metabolism plant Crassula argentea was substantially desensitized to the effects of regulatory ligands by treatment with diethylpyrocarbonate, a reagent which selectively modifies histidyl residues. Desensitization of the enzyme to the inhibitor malate and the activator glucose 6-phosphate was accompanied by the appearance of a peak in the ultraviolet difference spectrum at 240 nanometers, indicating the formation of ethoxyformylhistidyl derivatives. Hydroxylamine reversed part of the spectral change under native conditions, and almost all of the change under denaturing conditions, but failed to restore sensitivity to effectors. The pH profiles of desensitization to malate and glucose 6-phosphate indicated the involvement of groups on the enzyme with pK, values of 6.8 and 6.4, respectively. Under denaturing conditions, a total of 15 histidine residues per subunit were modified by diethylpyrocarbonate, whereas for the native enzyme nine histidines were modified per subunit. Effector desensitization occurs after the modification of two to three histidyl residues per subunit. The presence of malate reduced the apparent rate constant for desensitization by 60%, suggesting that the modification occurred at the malate binding site. Diethylpyrocarbonate treatment also eliminated the kinetic lag caused by malate. Glucose 6-phosphate did not protect the enzyme against diethylpyrocarbonate-induced desensitization.  相似文献   

17.
Modification of histidine residue(s) by diethylpyrocarbonate treatment of submitochondrial particles obtained by sonication results in inhibition of ATPase activity and stimulation of oligomycin-sensitive H+ conduction. The inhibition of the ATPase (EC 3.6.1.3) activity persisted in F1 isolated from diethylpyrocarbonate-treated submitochondrial particles, which exhibited the absorbance spectrum of modified histidine. Thus the inhibition of the ATPase activity results from histidine modification in F1 subunits. Removal of the natural inhibitor protein from submitochondrial particles resulted in stimulation of proton conduction. After removal of F1 inhibitor protein from the particles the stimulatory effect exerted by diethylpyrocarbonate treatment on proton conduction was lost. Reconstitution experiments showed that purified F1 inhibitor protein lost, after histidine modification, its capacity to inhibit the ATPase activity and proton conduction. These observations show that the stimulation of proton conduction by the ATPase complex effected by diethylpyrocarbonate treatment results from histidine modification in F1 inhibitor protein.  相似文献   

18.
Several cytochromes c2 from the Rhodospirillaceae show a pH dependence of redox potential in the physiological pH range which can be described by equations involving an ionisation in the oxidised form (pKo) and one in the reduced form (pKr). These cytochromes fall into one of two groups according to the degree of separation of pKo and pKr. In group A, represented here by the Rhodomicrobium vannielii cytochrome c2, the separation is approx. one pH unit and the ionisation is that of a haem propionic acid. Members of this group are unique among both cytochromes c2 and mitochondrial cytochromes c in lacking the conserved residue Arg-38. We propose that the role of Arg-38 is to lower the pK of the nearby propionic acid, so that it lies out of the physiological pH range. Substitution of this residue by an uncharged amino acid leads to a raised pK for the propionic acid. In group B, represented here by Rhodopseudomonas viridis cytochrome c2, the separation between pKo and pKr is approx. 0.4 pH unit and the ionisable group is a histidine at position 39. This was established by NMR spectroscopy and confirmed by chemical modification. Only a few other members of the cytochrome c2/mitochondrial cytochrome c family have a histidine at this position and of these, both Crithidia cytochrome c-557 and yeast cytochrome c were found to have a pH-dependent redox potential similar to that of Rps. viridis cytochrome c2. Using Coulomb's law, it was found that the energy required to separate pKo and pKr could be accounted for by simple electrostatic interactions between the haem iron and the ionisable group.  相似文献   

19.
The role of αPhe-291 residue in phosphate binding by Escherichia coli F1F0-ATP synthase was examined. X-ray structures of bovine mitochondrial enzyme suggest that this residue resides in close proximity to the conserved βR246 residue. Herein, we show that mutations αF291D and αF291E in E. coli reduce the ATPase activity of F1F0 membranes by 350-fold. Yet, significant oxidative phosphorylation activity is retained. In contrast to wild-type, ATPase activities of mutants were not inhibited by MgADP-azide, MgADP-fluoroaluminate, or MgADP-fluoroscandium. Whereas, 7-chloro-4-nitrobenzo-2-oxa-1,3-diazole (NBD-Cl) inhibited wild-type ATPase essentially completely, ATPase in mutants was inhibited maximally by ∼75%, although reaction still occurred at residue βTyr-297, proximal to αPhe-291 in the phosphate-binding pocket. Inhibition characteristics supported the conclusion that NBD-Cl reacts in βE (empty) catalytic sites, as shown previously by X-ray structure analysis. Phosphate protected against NBD-Cl inhibition in wild-type but not in mutants. In addition, our data suggest that the interaction of αPhe-291 with phosphate during ATP hydrolysis or synthesis may be distinct.  相似文献   

20.
Lin W 《Plant physiology》1979,63(5):952-955
Evidence is presented that K+ uptake in corn root segments is coupled to an electrogenic H+/K+ -exchanging plasmalemma ATPase while phosphate uptake is coupled to an OH/Pi antiporter. The plasmalemma ATPase inhibitor, diethylstilbestrol, or the stimulator, fusicoccin, altered K+ uptake directly and phosphate uptake indirectly. On the other hand, mersalyl, an OH/Pi antiporter inhibitor, inhibited phosphate uptake instantly but only slightly affected K+ uptake. Collapse of the proton gradient across the membrane by (p-trifluoromethoxy) carbonyl cyanide phenylhydrazone resulted in immediate inhibition of K+ uptake but only later inhibited phosphate uptake. Changing the pH of the absorption solution had opposite effects on K+ and phosphate uptake. In addition, a 4-hour washing of corn root tissue induced a 5-fold increase in the rate of K+ uptake with little or no lag, but only a 2- to 3-fold increase in phosphate uptake with a 30- to 45-minute lag. Collectively these differences strongly support the coupling of an electrogenic H+/K+ -exchanging ATPase to an OH/Pi antiporter in corn root tissue.  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