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 46 毫秒
1.
The kinetics of acetylation of mitochondrial succinate dehydrogenase [EC 1.3.99.1] in the two fibre types (A and C) of rat gastrocnemius with N-acetylimidazole was studied by a newly modified histochemical technique. Acetylimidazole partially inactivated the enzyme, but subsequent deacetylation with hydroxylamine restored the enzyme activity completely. Inactivation of the enzyme by acetylimidazole was prevented by malonate, which is a competitive inhibitor of the enzyme. The value of the inhibition constant (Ki = 34 microM) for malonate, obtained from the dependence of the pseudo-first order rate constant of acetylation of the enzyme with acetylimidazole on the malonate concentration, was in good agreement with the Ki value (33 microM) obtained by a different method, the dependence of the initial velocity of succinate oxidation by the dehydrogenase on the substrate concentration in the presence of malonate. These findings suggest that a tyrosyl residue is located in the malonate binding site (the active site) of succinate dehydrogenase in the gastrocnemius and plays a role in substrate binding, but is not a catalytic group.  相似文献   

2.
Cadmium ions inhibit membrane-bound succinate dehydrogenase with a second-order rate constant of 10.42 mM–1s–1 at pH 7.35 and 25°C. Succinate and malonate protect the enzyme against cadmium ion inhibition. The protection pattern exerted by succinate and malonate suggests that the group modified by cadmium is located at the active site. The pH curve of inactivation by Cd2+ indicates the involvement of an amino acid residue with pKa of 7.23.  相似文献   

3.
The rates of the oxidized (Eox) and reduced (Ered) (by NAD . H through the ubiquinone pool) succinate dehydrogenase inhibition by N-ethyl-maleimide are equal and obey pseudo-first order kinetics. The protection of the enzyme against irreversible alkylation was used to quantitate the dissociation constants for Eox and Ered complexes with fumarate, succinate and malonate under conditions when no intramolecular redox reactions might occur. the membrane-bound succinate dehydrogenase catalyzes the succinate : phenazine-methosulphate reductase reaction in the presence of thenoyltrifluoroacetone by a Slater-Bonner mechanism. A comparison of the constants measured by the protection with those derived from the steady-state kinetics shows that succinate affinity for Eox is about 10 times higher than that for Ered; the reverse relations were found for fumarate, whereas the affinity for malonate only slightly depends on the redox state of the enzyme. The data obtained suggest that the dicarboxylate binding at the active site induces changes in the enzyme redox potential. The surface charge does not contribute significantly to the energy of the dicarboxylate binding to the active site of the membrane-bound enzyme.  相似文献   

4.
Phenylglyoxal and 2,3-butanedione rapidly inactivate membrane-bound or soluble bovine heart succinate dehydrogenase. The inhibition of the enzyme by these reagents is completely prevented by saturating concentration of malonate. The modification of the active site sulfhydryl group by p-chloromercuribenzoate decreases the rate of the enzyme inhibition by phenylglyoxal and abolishes the protective effect of malonate. Kinetic data suggest that the inactivation by phenylglyoxal results from the modification of an essential arginine residue(s) which interacts with dicarboxylate to form the primary enzyme-substrate complex.  相似文献   

5.
Soluble succinate dehydrogenase prepared by butanol extraction reacts with N-ethylmaleimide according to first-order kinetics with respect to both remaining active enzyme and the inhibitor concentration. Binding of the sulfhydryl groups of the enzyme prevents its alkylation by N-ethylmaleimide and inhibition by oxaloacetate. A kinetic analysis of the inactivation of alkylating reagent in the presence of succinate or malonate suggests that N-ethylmaleimide acts as a site-directed inhibitor. The apparent first-order rate constant of alkylation increases between pH 5.8 and 7.8 indicating a pKa value for the enzyme sulfhydryl group equal to 7.0 at 22 degrees C in 50 mM Tris-sufate buffer. Certain anions (phosphate, citrate, maleate and acetate) decrease the reactivity of the enzyme towards the alkylating reagent. Succinate/phenazine methosulfate reductase activity measured in the presence of a saturating concentration of succinate shows the same pH-dependence as the alkylation rate by N-ethylmaleimide. The mechanism of the first step of succinate oxidation, including a nucleophilic attack of substrate by the active-site sulfhydryl group, is discussed.  相似文献   

6.
The effect of disulfiram on succinate oxidase and succinate dehydrogenase activities of beef heart submitochondrial particles was studied. Results show that disulfiram inhibits both functions. Succinate and malonate suppress the inhibitory action of disulfiram when succinate dehydrogenase is stabilized in an active conformation. Disulfiram is not able to inhibit the enzyme when succinate dehydrogenase is inactivated by oxaloacetate. The inhibitory effect of disulfiram is reverted by the addition of dithiothreitol. From these results, it is proposed that disulfiram inhibits the utilization of succinate by a direct modification of an -SH group located in the catalytically active site of succinate dehydrogenase.  相似文献   

7.
(1) Purified succinate dehydrogenase contains about 49 mol of lysine residues per mol enzyme. Titration of succinate dehydrogenase with fluorescamine indicates that half the lysyl groups are located on the surface of the protein and the other half are buried inside. (2) The reconstitutive activity and the low Km ferricyanide reductase activity of succinate dehydrogenase decreased as the extent of alkylation of amino groups by fluorescamine increased. (3) The inhibitory effects of fluorescamine on both activities are parallel and are succinate concentration dependent. (4) Alkylation of the native succinate-Q reductase by fluorescamine does not affect the enzymatic activity or alter the enzyme kinetic parameters. This indicates that the inhibitory effect of fluorescamine on succinate dehydrogenase is due to the modification of a specific amino group(s) on succinate dehydrogenase which is essential in the interaction with QPs to form succinate-Q reductase. The participation of an ionic group in the formation of succinate-Q reductase supports the idea of the involvement of ionic interaction between succinate dehydrogenase and QPs.  相似文献   

8.
The effect of structural analogues of l-malate was studied on NADP-malic enzyme purified from Zea mays L. leaves. Among the compounds tested, the organic acids behaved as more potent inhibitors at pH 7.0 than at pH 8.0, suggesting that the dimeric form was more susceptible to the inhibition than the tetrameric form of the enzyme.Oxalate, ketomalonate, hydroxymalonate, malonate, oxaloacetate, tartrate, -hydroxybutyrate, -ketobutyrate, -ketoglutarate and -hydroxyglutarate exhibited linear competitive inhibition with respect to the substrate l-malate at pH 8.0. On the other hand, glyoxylate and glycolate turned out to be non-competitive inhibitors, while glycolaldehyde, succinate, fumarate, maleate and - and -hydroxybutyrate had no effect on the enzyme activity, at the concentrations assayed. These results suggest that the extent of inhibition was dependent on the size of the analogues and that the presence of an 1-carboxyl group along with a 2-hydroxyl or 2-keto group was important for binding of the substrate analogue to the enzyme.  相似文献   

9.
Abstract: Although the mechanism of neuronal death in neurodegenerative diseases remains unknown, it has been hypothesized that relatively minor metabolic defects may predispose neurons to N -methyl- d -aspartate (NMDA) receptor-mediated excitotoxic damage in these disorders. To further investigate this possibility, we have characterized the excitotoxic potential of the reversible succinate dehydrogenase (SDH) inhibitor malonate. After its intrastriatal stereotaxic injection into male Sprague-Dawley rats, malonate produced a dose-dependent lesion when assessed 3 days after surgery using cytochrome oxidase histochemistry. This lesion was attenuated by coadministration of excess succinate, indicating that it was caused by specific inhibition of SDH. The lesion was also prevented by administration of the noncompetitive NMDA antagonist MK-801. MK-801 did not induce hypothermia, and hypothermia itself was not neuroprotective, suggesting that the neuroprotective effect of MK-801 was due to blockade of the NMDA receptor ion channel and not to any nonspecific effect. The competitive NMDA antagonist LY274614 and the glycine site antagonist 7-chlorokynurenate also profoundly attenuated malonate neurotoxicity, further indicating an NMDA receptor-mediated event. Finally, the α-amino-3-hydroxy-5-methyl-4-isoxazolepropionate (AMPA) antagonist NBQX (2,3-dihydroxy-6-nitro-7-sulfamoylbenzo( f )-quinoxaline) was ineffective at preventing malonate toxicity at a dose that effectively reduced S -AMPA toxicity, indicating that non-NMDA receptors are involved minimally, if at all, in the production of the malonate lesion. We conclude that inhibition of SDH by malonate results in NMDA receptor-mediated excitotoxic neuronal death. If this mechanism of "secondary" or "weak" excitotoxicity plays a role in neurodegenerative disease, NMDA antagonists and other "antiexcitotoxic" strategies may have therapeutic potential for these diseases.  相似文献   

10.
The enzyme Δ1-pyrroline-5-carboxylate (P5C) dehydrogenase (aka P5CDH and ALDH4A1) is an aldehyde dehydrogenase that catalyzes the oxidation of γ-glutamate semialdehyde to l-glutamate. The crystal structures of mouse P5CDH complexed with glutarate, succinate, malonate, glyoxylate, and acetate are reported. The structures are used to build a structure-activity relationship that describes the semialdehyde carbon chain length and the position of the aldehyde group in relation to the cysteine nucleophile and oxyanion hole. Efficient 4- and 5-carbon substrates share the common feature of being long enough to span the distance between the anchor loop at the bottom of the active site and the oxyanion hole at the top of the active site. The inactive 2- and 3-carbon semialdehydes bind the anchor loop but are too short to reach the oxyanion hole. Inhibition of P5CDH by glyoxylate, malonate, succinate, glutarate, and l-glutamate is also examined. The Ki values are 0.27 mM for glyoxylate, 58 mM for succinate, 30 mM for glutarate, and 12 mM for l-glutamate. Curiously, malonate is not an inhibitor. The trends in Ki likely reflect a trade-off between the penalty for desolvating the carboxylates of the free inhibitor and the number of compensating hydrogen bonds formed in the enzyme-inhibitor complex.  相似文献   

11.
A sterile glucose-mineral salts broth was fortified with equimolar concentrations (10-3 M) of various organic acids and intermediates in the tricarboxylic acid cycle. Appropriate media were neutralized with 2 N NaOH, inoculated with spore suspensions or mycelial pellets ofPenicillium rubrum and incubated quiescently for 14 days or with shaking for 5 days. Rubratoxins were recovered from culture filtrates by ether extraction and resolved by thin-layer chromatography. Toxin formation in quiescent cultures was enhanced by malonate but was not markedly affected by ethyl malonate, shikimate, and acetate or by isocitrate or oxaloacetate added in the presence of malonate. Citrate, cis-aconitate, -ketoglutarate, succinate, fumarate, and malonate when present in the medium alone or in conjunction with malonate caused a 15 to 50% reduction in rubratoxin formation. Acetyl-CoA (10-5 M/flask) caused an 80% increase in toxin yield. Rubratoxin formation in shake cultures was not affected by succinate and malonate. All other combinations of intermediates and malonate caused a 10 to 50% reduction in toxin formation. At 10–3 M, citrate enhanced rubratoxin B formation and stimulated rubratoxin A production by as much as 100%. Above 10–3 M, citrate inhibited toxin production. Incorporation of [2-14C]acetate into rubratoxin was enhanced by malonate, fumarate, and malonate. A combination of pyruvate and malonate produced a 40% increase in [2-14C]acetate incorporation into rubratoxin. The highest reduction of labeled acetate incorporation (36%) was caused by succinate or -ketoglutarate combined with malonate.  相似文献   

12.
Purified and membrane-bound succinate dehydrogenase (SDH) from bovine heart mitochondria was inhibited by the histidine-modifying reagents ethoxyformic anhydride (EFA) and Rose Bengal in the presence of light. Succinate and competitive inhibitors protected against inhibition, and decreased the number of histidyl residues modified by EFA. The essential residue modified by EFA was not the essential thiol of SDH, but modification of the essential thiol abolished the protective effect of malonate against inhibition of SDH by EFA. The EFA inhibition was reversed by hydroxylamine nearly completely when the inhibition was less than or equal to 35%, and only partially when the inhibition was more extensive. The uv spectrum of EFA-modified SDH before and after hydroxylamine treatment suggested that extensive inhibition of SDH with EFA may result in ethoxyformylation at both imidazole nitrogens of histidyl residues. Such a modification is not reversed by hydroxylamine. Succinate dehydrogenases and fumarate reductases from several different sources have similar compositions, and the two enzymes from Escherichia coli have considerable homology in the amino acid composition of their respective flavoprotein and iron-sulfur protein subunits. In the former, there is a short stretch containing conserved histidine, cysteine, and arginine residues. These residues, if also conserved in the bovine enzyme, may be the essential active site residues suggested by this work (histidine) and previously (cysteine, arginine).  相似文献   

13.
14.
Respiratory complex II (CII, succinate dehydrogenase, SDH) inhibition can induce cell death, but the mechanistic details need clarification. To elucidate the role of reactive oxygen species (ROS) formation upon the ubiquinone-binding (Qp) site blockade, we substituted CII subunit C (SDHC) residues lining the Qp site by site-directed mutagenesis. Cell lines carrying these mutations were characterized on the bases of CII activity and exposed to Qp site inhibitors MitoVES, thenoyltrifluoroacetone (TTFA) and Atpenin A5. We found that I56F and S68A SDHC variants, which support succinate-mediated respiration and maintain low intracellular succinate, were less efficiently inhibited by MitoVES than the wild-type (WT) variant. Importantly, associated ROS generation and cell death induction was also impaired, and cell death in the WT cells was malonate and catalase sensitive. In contrast, the S68A variant was much more susceptible to TTFA inhibition than the I56F variant or the WT CII, which was again reflected by enhanced ROS formation and increased malonate- and catalase-sensitive cell death induction. The R72C variant that accumulates intracellular succinate due to compromised CII activity was resistant to MitoVES and TTFA treatment and did not increase ROS, even though TTFA efficiently generated ROS at low succinate in mitochondria isolated from R72C cells. Similarly, the high-affinity Qp site inhibitor Atpenin A5 rapidly increased intracellular succinate in WT cells but did not induce ROS or cell death, unlike MitoVES and TTFA that upregulated succinate only moderately. These results demonstrate that cell death initiation upon CII inhibition depends on ROS and that the extent of cell death correlates with the potency of inhibition at the Qp site unless intracellular succinate is high. In addition, this validates the Qp site of CII as a target for cell death induction with relevance to cancer therapy.Mitochondrial respiratory complex II (CII), aka succinate dehydrogenase (SDH), directly links the tricarboxylic acid (TCA) cycle to the electron transport chain (ETC) by mediating electron transfer from the TCA cycle metabolite succinate to ubiquinone (UbQ).1 For this reason, CII is subjected to a high electron flux between the succinate-binding dicarboxylate site in the matrix-exposed subunit A and the proximal UbQ-binding (Qp) site, formed by the subunits C (SDHC) and D embedded in the mitochondrial inner membrane (Figure 1b).2, 3, 4, 5 Disruption of electron transfer to UbQ, for example by Qp site inhibition, leads to reactive oxygen species (ROS) generation from CII due to the leakage of ‘stalled'' electrons to molecular oxygen at the reduced flavin adenine dinucleotide (FAD) prosthetic group. However, ROS production from reduced FAD is only possible when the adjacent dicarboxylate site is neither occupied by its substrate succinate, typically at low succinate conditions, nor inhibited by other dicarboxylates, for example by malonate.6, 7, 8, 9, 10Open in a separate windowFigure 1Amino-acid substitutions in the Qp site of CII. (a) Multiple species alignment of the SDHC region bordering the Qp site shows a high level of conservation. Amino-acid substitutions prepared for this study are indicated in human SDHC. (b) Three dimensional representation of CII and the topology of the Qp site. SDHC residues mutated in this study are indicated by arrows. Displayed is the humanized crystal structure of porcine CII.3 (c) A snapshot from molecular dynamics simulation of MitoVES interaction with the Qp site of CII in the presence of phospholipid bilayer.16 One of the possible conformations of MitoVES is shown in orange, substituted SDHC residues are depicted in magentaBeyond bioenergetics, CII has emerged as an important factor in cell death induction.11, 12 On one hand, it has been proposed that increased ROS production from CII, resulting from changes in matrix pH and calcium status, amplifies cell death signals originating at other sites.12, 13, 14, 15 On the other hand, the inhibition of CII may also directly initiate cell death, as suggested by our previous results with vitamin E (VE) analogs such as the mitochondrially targeted VE succinate (MitoVES). This compound inhibits CII activity leading to ROS generation and cell death induction in cancer cells, as evidenced by the suppression of tumor growth in experimental animal models.16, 17, 18, 19, 20 The efficacy of MitoVES is greatly reduced in the absence of functional CII, and computer modeling along with other corroborative evidence suggests that MitoVES binds to the Qp site of CII.16 However, this is only circumstantial evidence with respect to cell death induction, as cells lacking electron flux within CII due to a structural defect should not be able to produce CII-derived ROS. Accordingly, not only the direct cell death initiation upon CII inhibition will be compromised in this situation, but also the indirect signal amplification mentioned above will be affected.In the present study, we combined site-directed mutagenesis of Qp site amino-acid residues with the use of Qp site inhibitors MitoVES, thenoyltrifluoroacetone (TTFA) and Atpenin A5 to assess the link between Qp site inhibition and cell death initiation. We show that for MitoVES and TTFA, the potency of Qp site inhibition correlates with the extent of ROS production and cell death induction in respiration-competent CII variants, and that the induced cell death is dependent on CII-derived ROS.Atpenin, however, did not induce cell death, possibly due to the rapid accumulation of succinate in intact cells, incompatible with ROS generation from CII. These results provide evidence for the role of CII in cell death initiation and establish the Qp site as a target for cell death induction.  相似文献   

15.
Summary The kinetic behaviour of succinate dehydrogenase [EC 1.3.99.1] in three fibre types of rat gastrocnemius was examined by a quantitative histochemical method without disruption of the cellular structure. 2-(2-Benzothiazolyl)-3-(4-phthalhydrazidyl)-5-styryl-tetrazolium chloride (BPST) and phenazine methosulphate were used as electron acceptors. On measurement of the absorbance value at 530 nm of BPST formazan, produced by the succinate dehydrogenase reaction in sections, it was found that the staining intensity of succinate dehydrogenase was linearly proportional to both the incubation time and the thickness of the slice therefore, the initial velocity of the staining could be calculated. By Michaelis-Menten (1913) treatment of the dependence of the initial velocity on the substrate concentration in the absence and the presence of a competitive inhibitor, malonate, the Km andVmax values for succinate and the Ki value for malonate were obtained. The Km and Ki values of the three fibre types were similar. The ratio of theVmax values of type A, B and C fibres was 1.02.03.3. The temperature dependence of the kinetic parameters was very similar in the three fibre types. These findings confirm that the differences in the staining intensity of the three fibre types reflect differences in the amounts, but not the properties, of succinate dehydrogenase.  相似文献   

16.
Summary The mitochondrial succinate dehydrogenase (E.C. 1.3.3.99) is subjected to apparently complicated regulatory mechanism. Yet, systematic analysis of the mechanism reveals the simplicity of the control. There are two stable forms of the enzyme; the non-active form stabilized as 1:1 complex with oxaloacetate and the active form stabilized by binding of activating ligands. This model quantitatively describes either the equilibrium level of active enzyme or the kinetics of activation-deactivation, in the presence of various concentrations of opposing effectors. The site where the regulatory ligands interact with the enzyme is not the substrate bonding site. The marked differences of dissociation constants of the same ligand from the two sites clearly distinguish between them.This model is fully developed for simple cases where the activating ligands are dicarboxylic acids or monovalent anions. On the other hand with activators such as ATP or CoQH2, quantitation is still not at hand. This stems from the difficulties in maintaining determined, measurable, concentrations of the ligand in equilibrium with the membranal enzyme.While in active form the histidyl flavin moity of the enzyme is reduced by physiological substrate (succinate; CoQH2). The non-active form is not reduced by these compounds, only strong reductants with low redox potential reduce the non-active enzyme. It is suggested that deactivation is a simple modulation of the redox potential of the flavin form E 0 mV in the active enzyme to E < –190 mV. The switch from one state to another might be achieved by distortion of the planar form of oxidized flavin to the bend configuration of the reduced flavin. Thus, in the active enzyme such distortion will destabilize the oxidized state of the flavin, shifting the redox potential to the higher value. The binding of oxaloacetate to the regulatory sites releases the distorting forces by relaxing the conformation of the enzyme. Consequently, the flavin assumes its planar form with the low redox potential. This assumption is supported by the spectral shifts of the flavin associated with the activation deactivation transition.The suicidal oxidation of malate to oxaloacetate, carried by the succinate dehydrogenase, plays an important role in modulating the enzyme activity in the mitochondria. This mechanism might supply oxaloacetate for deactivation in spite of the negligible concentration of free oxaloacetate in the matrix. The oxidation of malate by the enzyme is controlled by the redox potential at the immediate vicinity of the enzyme, and is imposed by the redox level of the membranal quinone.Finally, the modulation of succinate dehydrogenase activity is closely associated with regulation of NADH oxidation through the mutual inhibition between oxidases (Gutman, M. in Bioenergetics of Membranes, L. Packer et al., ed. Elsevier 1977, p. 165). The consequence of these interactions is the selection for the main electron donnor for the respiratory chain, during mixed substrate respiration, according to the metabolic demands from the mitochondria.Abbreviations SDH succinate dehydrogenase (succinate: acceptor oxidoreductase (E.C. 1.3.99.1)); - OAA oxaloacetate - Act activator - EA, EA active and non active forms of the enzyme, respectively - K'eq apparent equilibrium constant - K'd apparent dissociation constant - KAct, KOAA dissociation constant of the respective ligand from the enzyme - K'a, k'd the apparent rate constants of activation and deactivation, respectively - ka, kd the true rate constant of activation and deactivation respectively - ETP, ETPII non phosphorylating and phosphorylating submitochondrial particles - PMS phenazine methosulfate - DCIP dichlorophenol indophenol - CoQ ubiquinone - TIFA Thenotriflouvoacetone - NEM N methyl Maleimide  相似文献   

17.
The effect of malonate and sedaxane, a compound with the fungicidal effect which act as succinate dehydrogenase inhibitors, on the resistance of etiolated wheat seedlings (Triticum aestivum L.) to osmotic stress caused by 12% PEG 6000 solution, was studied. The presowing treatment of seeds with 0.3 mM sedaxane solution significantly reduced the inhibitory effect of osmotic stress on seedling growth. The protective effect of 10 mM malonate was significant when it was added to the incubation medium of the roots; the effect of preseeding treatment with malonate was less significant. Unlike malonate, malate had no positive effect on seedling growth under osmotic stress. The activity of succinate dehydrogenase and the hydrogen peroxide content decreased in seedlings after the treatment of roots with malonate and sedaxane. Pretreatment with sedaxane and the addition of malonate to the incubation medium of roots prevented the accumulation of a lipid peroxidation product, malondialdehyde, which is caused by osmotic stress, and increased peroxidase activity. It was concluded that the stress-protective effect of sedaxane and malonate on wheat seedlings might be due to the inhibition of succinate dehydrogenase-dependent formation of reactive oxygen species and the prevention of oxidative cell damage.  相似文献   

18.
The effect of cadmium (Cd2+), mercury (Hg2+) and copper (Cu2+) was studied with partially purified flavokinase (ATP:riboflavin 5-phosphotransferase EC 2.7.1.26) from rat liver. All the divalent heavy metal cations inhibited flavokinase activity in a concentration-dependent manner. The inhibitory effect of cadmium on the enzyme was completely reversed by increasing concentration, of Zinc (Zn2+) indicating a competition between Zn2+ and Cd2+ for binding with the enzyme. A competition between riboflavin and Cd2+ is also evident from the present investigation. These observations hint at the possibility that Zn2+ and Cd2+ probably compete for the same site on the enzyme where riboflavin binds. However, inhibition of flavokinase by Hg2+ could not be reversed by Zn2+. Our studies further reveal that hepatic flavokinase appears to contain an essential, accessible and functional thiol group(s) which is evident from a concentration dependent inhibition of activity by sulfhydryl reagent s like parachloromercuribenzoate (PCMB), 5,5-dithiobis (2-nitrobenzoic acid)(DTNB), and N-ethylmaleimide (NEM). Inhibition of flavokinase by sulfhydryl reagents were protected, except in case of NEM inhibition, when the enzyme was incubated with thiol protectors like glutathione (GSH) and dithiothreitol (DTT). Furthermore, the enzyme could also be protected from the inhibitory effect of Cd2+ and Hg2+ by GSH and DTT suggesting that Cd2+ probably interacts with a reactive thiol group at or near the active site of enzyme in bringing about its inhibitory effect. (Mol Cell Biochem 167: 73-80, 1997)  相似文献   

19.
Increased levels of maleate cis-trans isomerase in Alcaligenes faecalis IB-14 were induced by malonate analogues such as tartronate, keto-malonate and ethyl-malonate. Similarly in the case of malonate, those analogues were not served as carbon sources for the cell growth, but they were more effective as inducer than maleate, the normal inducer of the enzyme.

The induced synthesis of the enzyme was markedly repressed in the presence of various carbon sources, e.g. oxalacetate, d- and l-malates, fumarate and succinate. More than 90 per cent of the enzyme formation was inhibited when 10?1 m of oxalacetate or dl-malate was present in malonate medium. The repression of the enzyme synthesis by those dicarboxylic acids was always associated with remarkable promotion of the cell growth.  相似文献   

20.
Malonyl-CoA synthetase fromPseudomonas fluorescens was inactivated by diethylpyrocarbonate (DEP) with the second-order rate constant of 775 M–1 min–1 atpH 7.0, 25°C, showing a concomitant increase in absorbance at 242 nm due to the formation of N-carbethoxyhistidyl derivatives. The inactivated enzyme at low concentration of DEP (<0.2 mM) could be completely reactivated by hydroxylamine but not completely reactivated at high concentration (>0.5 mM), indicating that there may be another functional group modified by DEP. Complete inactivation of malonyl-CoA synthetase required the modification of seven residues per molecule of enzyme; however, only one is calculated to be essential for enzyme activity by a statistical analysis of the residual enzyme activity.pH dependence of inactivation indicated the involvement of a residue with apK a of 6.7, which is closely related to that of histidyl residue of proteins. Whena subunit treated with DEP was mixed with subunits complex, the enzyme activity completely disappeared, whereas when subunit complex treated with the reagent was mixed witha subunit, the activity remained. Inactivation of the enzyme by the reagent was protected by the presence of malonate and ATP. These results indicate that a catalytically essential histidyl residue is located at or near the malonate and ATP binding region ona subunit of the enzyme.  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