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.
The reaction mechanism of aspartate transcarbamylase from mouse spleen has been determined, using steady-state kinetics, isotope-exchange experiments, inhibition studies with a transition-state analog, and product-inhibition studies. Intersecting reciprocal plots obtained when one substrate was varied against different concentrations of the second substrate indicate that the mechanism is sequential. The transition-state analog, N-(phosphonacetyl)-l-aspartate, was a powerful inhibitor of aspartate transcarbamylase, with an inhibition constant (Ki) of 2.6 × 10?8m at 37 °C and pH 7.4 in 0.05 m Na HEPES buffer. PALA gave competitive inhibition with carbamyl phosphate and noncompetitive inhibition with l-aspartate, indicating that carbamyl phosphate must bind before aspartate for catalysis to occur. A ping-pong mechanism in which carbamyl phosphate binds first was excluded by isotope-exchange experiments, since [32P]inorganic phosphate was not incorporated into carbamyl phosphate in the absence of aspartate. Product-inhibition studies showed that only inorganic phosphate and carbamyl phosphate gave a competitive pattern; all other combinations of substrate and product gave noncompetitive inhibition patterns when incubations were carried out at subsaturating concentrations of the second substrate. These inhibition patterns showed that carbamyl phosphate binds first, aspartate binds second, carbamyl aspartate dissociates first, and phosphate dissociates second.  相似文献   

2.
N-Methylpurine-DNA glycosylase (MPG) initiates base excision repair in DNA by removing a wide variety of alkylated, deaminated, and lipid peroxidation-induced purine adducts. In this study, we tested the role of excised base on MPG enzymatic activity. After the reaction, MPG produced two products: free damaged base and AP-site containing DNA. Our results showed that MPG excises 1,N6-ethenoadenine (?A) from ?A-containing oligonucleotide (?A-DNA) at a similar or slightly increased efficiency than it does hypoxanthine (Hx) from Hx-containing oligonucleotide (Hx-DNA) under similar conditions. Real-time binding experiments by surface plasmon resonance (SPR) spectroscopy suggested that both the substrate DNAs have a similar equilibrium binding constant (KD) towards MPG, but under single-turnover (STO) condition there is apparently no effect on catalytic chemistry; however, the turnover of the enzyme under multiple-turnover (MTO) condition is higher for ?A-DNA than it is for Hx-DNA. Real-time binding experiments by SPR spectroscopy further showed that the dissociation of MPG from its product, AP-site containing DNA, is faster than the overall turnover of either Hx- or ?A-DNA reaction. We thereby conclude that the excised base plays a critical role in product inhibition and, hence, is essential for MPG glycosylase activity. Thus, the results provide the first evidence that the excised base rather than AP-site could be rate-limiting for DNA-glycosylase reactions.  相似文献   

3.
A mathematical model for enzymatic cellulose hydrolysis, based on experimental kinetics of the process catalysed by a cellulase [see 1,4-(1,3;1,4)-β-d-glucan 4-glucanohydrolase, EC 3.2.1.4] preparation from Trichoderma longibrachiatum has been developed. The model takes into account the composition of the cellulase complex, the structural complexity of cellulose, the inhibition by reaction products, the inactivation of enzymes in the course of the enzymatic hydrolysis and describes the kinetics of d-glucose and cellobiose formation from cellulose. The rate of d-glucose formation decelerated through the hydrolysis due to a change in cellulose reactivity and inhibition by the reaction product, d-glucose. The rate of cellobiose formation decelerated due to inhibition by the product, cellobiose, and inactivation of enzymes adsorbed on the cellulose surface. Inactivation of the cellobiose-producing enzymes as a result of their adsorption was found to be reversible. The model satisfactorily predicts the kinetics of d-glucose and cellobiose accumulation in a batch reactor up to 70–80% substrate conversion on changing substrate concentration from 5 to 100 g l?1and the concentration of the enzymic preparation from 5 to 60 g l?1.  相似文献   

4.
Sirtuins are key regulators of many cellular functions including cell growth, apoptosis, metabolism, and genetic control of age-related diseases. Sirtuins are themselves regulated by their cofactor nicotinamide adenine dinucleotide (NAD+) as well as their reaction product nicotinamide (NAM), the physiological concentrations of which vary during the process of aging. Nicotinamide inhibits sirtuins through the so-called base exchange pathway, wherein rebinding of the reaction product to the enzyme accelerates the reverse reaction. We investigated the mechanism of nicotinamide inhibition of human SIRT3, the major mitochondrial sirtuin deacetylase, in vitro and in silico using experimental kinetic analysis and Molecular Mechanics-Poisson Boltzmann/Generalized Born Surface Area (MM-PB(GB)SA) binding affinity calculations with molecular dynamics sampling. Through experimental kinetic studies, we demonstrate that NAM inhibition of SIRT3 involves apparent competition between the inhibitor and the enzyme cofactor NAD+, contrary to the traditional characterization of base exchange as noncompetitive inhibition. We report a model for base exchange inhibition that relates such kinetic properties to physicochemical properties, including the free energies of enzyme-ligand binding, and estimate the latter through the first reported computational binding affinity calculations for SIRT3:NAD+, SIRT3:NAM, and analogous complexes for Sir2. The computational results support our kinetic model, establishing foundations for quantitative modeling of NAD+/NAM regulation of mammalian sirtuins during aging and the computational design of sirtuin activators that operate through alleviation of base exchange inhibition.  相似文献   

5.
The kinetic mechanism of action of Escherichia coli acetate kinase (ATP: acetate phosphotransferase, EC 2.7.2.1) was found to be Ping Pong. This conclusion was alluded to from initial rate experiments and from competitive and product inhibition experiments. The findings are consistent with chemical evidence providing the basis for the participation of a phosphoryl-enzyme intermediate in the reaction.  相似文献   

6.
Procedures are described for isolating highly purified porcine liver pyruvate and α-ketoglutarate dehydrogenase complexes. Rabbit serum stabilized these enzyme complexes in mitochondrial extracts, apparently by inhibiting lysosomal proteases. The complexes were purified by a three-step procedure involving fractionation with polyethylene glycol, pelleting through 12.5% sucrose, and a second fractionation under altered conditions with polyethylene glycol. Sedimentation equilibrium studies gave a molecular weight of 7.2 × 106 for the liver pyruvate dehydrogenase complex. Kinetic parameters are presented for the reaction catalyzed by the pyruvate dehydrogenase complex and for the regulatory reactions catalyzed by the pyruvate dehydrogenase kinase and pyruvate dehydrogenase phosphatase. For the overall catalytic reaction, the competitive Ki to Km ratio for NADH versus NAD+ and acetyl CoA versus CoA were 4.7 and 5.2, respectively. Near maximal stimulations of pyruvate dehydrogenase kinase by NADH and acetyl CoA were observed at NADH:NAD+ and acetyl CoA:CoA ratios of 0.15 and 0.5, respectively. The much lower ratios required for enhanced inactivation of the complex by pyruvate dehydrogenase kinase than for product inhibition indicate that the level of activity of the regulatory enzyme is not directly determined by the relative affinity of substrates and products of catalytic sites in the pyruvate dehydrogenase complex. In the pyruvate dehydrogenase kinase reaction, K+ and NH+4 decreased the Km for ATP and the competitive inhibition constants for ADP and (β,γ-methylene)adenosine triphosphate. Thiamine pyrophosphate strongly inhibited kinase activity. A high concentration of ADP did not alter the degree of inhibition by thiamine pyrophosphate nor did it increase the concentration of thiamine pyrophosphate required for half-maximal inhibition.  相似文献   

7.
The L-tert-leucine synthesis was performed continuously in series of two enzyme-membrane reactors by reductive amination of trimethylpyruvate with leucine dehydrogenase. The necessary “native” cofactor NADH is regenerated with the aid of a second enzyme, formate dehydrogenase. Considering detailed kinetic studies of initial reaction rates under conditions relevant to the process a kinetic model was developed. The model shows that the overall reaction rate is strongly inhibited by the reaction product. The reactor's models combine the mass balances and proposed kinetic equations. The model adequacy was verified by using it to simulate the experiments and by comparing experimental and computed conversion, space-time yield and enzyme consumption. The calculations for the three reactor's types (batch, single CSTR and a cascade of two CSTRs in series) were compared. The results showed that a single CSTR is no favourable reactor configuration due to the very strong product inhibition. Space-time yield drops from 560 g litre?1 day?1 in a batch reactor to 110 g litre?1 day?1 in a single CSTR at the highest conversion of 98%. At the conversion of 95% the difference in biocatalyst costs between batch and two CSTR in series is negligible. Therefore the use of two enzyme membrane reactors in series was proposed. The modelling in this work shows that the optimisation of the quantity of the enzyme used results in a minimisation of the biocatalyst costs.  相似文献   

8.
The reaction of NAD(P)H:flavin oxidoreductase (flavin reductase) from Photobacterium fischeri is proposed to follow a ping-pong bisubstrate-biproduct mechanism. This is based on a steady-state kinetic analysis of initial velocities and patterns of inhibition by NAD+ and AMP. The double reciprocal plots of initial velocities versus concentrations of FMN or NADH show, in both cases, a series of parallel lines. The Michaelis constants for NADH (FMN saturating) and FMN (NADH saturating) are 2.2 and 1.2 × 10?4m, respectively. The product NAD+ has been found to be an inhibitor competitive with FMN but non-competitive with NADH. Using AMP as an inhibitor, noncompetitive inhibition patterns were observed with respect to both NADH and FMN as the varied substrate. In addition, the reductase was not inactivated by treatment with N-ethylmaleimide either alone or in the presence of FMN, but the enzyme was inactivated by N-ethylmaleimide in the presence of NADH. These findings suggest that flavin reductase shuttles between disulfide- and sulfhydryl-containing forms during catalysis.  相似文献   

9.
10.
N-Nitrosodialkylamines are known to be potent indirect-acting mutagens/carcinogens, which are activated by cytochrome P450. The reaction product of N-nitroso-N-methylbutylamine (NMB) with modified Fenton’s reagent supplemented with copper salt (Fe2+–Cu2+–H2O2) was reported to be mutagenic in Salmonella typhimurium TA1535 without S9 mix. In this study, the NMB activation mechanism was investigated by ESR spectroscopy with radical trapping agents to detect radical species and also by observing changes in mutagenic potency with a Salmonella strain in the Ames assay in the presence of radical trapping agents. In ESR spectroscopy experiments, the hydroxyl radical generated from the modified Fenton’s reagent was detected using the hydroxyl radical trapping agent 5,5-dimethyl-1-pyrroline N-oxide (DMPO). Since the amount of the DMPO–OH adduct decreased with the addition of NMB, hydroxyl radical was presumed to react with NMB followed by the generation of nitric oxide (NO), which was detected as CarboxyPTI through reaction with 2-(4-carboxyphenyl)-4,4,5,5-tetramethylimidazoline-1-oxyl 3-oxide (CarboxyPTIO). The mutagenicity of the reaction extract decreased following the addition of DMPO or CarboxyPTIO. Furthermore, the mutagenicity of the reaction product in the presence of DMPO was enhanced by the addition of NO. The reaction product from NMB with Fe2+–Cu2+–NO in the absence of H2O2 was mutagenic, and this activity increased with the introduction of additional NO. These findings suggest that hydroxyl radical takes part in the generation of NO from NMB and that NO plays an important role in NMB activation in the presence of Fe2+ and Cu2+.  相似文献   

11.
Glyceraldehyde-3-phosphate dehydrogenase (GAPD) was isolated from human erythrocyte ghosts by a simple procedure utilizing ammonium sulfate precipitation and affinity chromatography on NAD+-Sepharose 4B. The purified enzyme had a specific activity of 98 units/mg protein. The kinetic mechanism of GAPD was studied by product and deadend inhibition using NADH, α-glycerophosphate, nitrate, and 2,3-diphosphoglycerate. The results indicated that the human erythrocyte GAPD-catalyzed reaction follows an ordered ter bi mechanism characterized by the sequential addition of NAD+, glyceraldehyde 3-phosphate (GAP), and phosphate to the enzyme and the sequential release of 1,3-diphosphoglycerate and NADH from the enzyme. This contrasts with the mechanism (rapid equilibrium random ter bi) proposed by Oguchi (1970, J. Biochem. (Tokyo)68, 427–439) who based his conclusion on the initial rate data alone. Since the Michaelis-Menten kinetics were not applicable to this enzyme because of the competitive substrate inhibition by GAP, we devised a new kinetic approach for determining the parameters of the GAPD-catalyzed reaction. Results of this study indicate that the GAPD-catalyzed reaction is regulated by both ATP and GAP. We propose that GAP acts as an “amplifier” for the feedback inhibition effect of ATP. We discuss the effect this may have played in causing controversy over the regulatory role of this enzyme in glycolysis.  相似文献   

12.
The effect of media supplements on total and polyglutamylfolate concentrations has been examined in Neurospora crassa wild type (FGSC 853), an ethionine-resistant mutant (FGSC 1212), and a methionine auxotroph (FGSC 1330) which lacks folylpolyglutamate synthetase. When the culture medium contained 1 mm glycine, folate concentrations in the wild type were increased by over 90% and more p-[3H]aminobenzoate was incorporated into folates. Growth in l-methionine-supplemented media (1–5 mm) decreased folate levels and labeling in all three strains. In the wild type, this effect of l-methionine was reversed on transfer to unsupplemented media but p-[3H]aminobenzoate pulse-chase experiments suggested that exogenous methionine did not increase the turnover of labeled folates. At 1 mm, d-methionine did not affect polyglutamylfolate labeling but l-methionine reduced 3H incorporation by 65% in the wild type. Ion-exchange chromatography showed that p-[3H]aminobenzoate was incorporated in formyl- and methyltetrahydrofolates which in the wild type, were principally hexaglutamyl derivatives. Glycine-supplemented growth yielded labeled folates that were 24% heptaglutamates but these and pentaglutamates were lacking when l-methionine was supplied. The specific activity of GTP cyclohydrolase was not significantly affected by culture in l-methionine-containing media. Dialysis and gel filtration both lowered enzyme activities and product formation was not changed when up to 10 μmol of l-methionine was added to the reaction system. The data suggest that methionine or its metabolic products exerts some control over folate production which is distinct from the established inhibition of methylenetetrahydrofolate reductase by AdoMet.  相似文献   

13.
Stopped-flow kinetic data were obtained for the iron-type nitrile hydratase from Rhodococcus equi TG328-2 (ReNHase) using methacrylonitrile as the substrate. Multiple turnover experiments suggest a three-step kinetic model that allows for the reversible binding of substrate, the presence of an intermediate, and the formation of product. Microscopic rate constants determined from these data are in good agreement with steady state data confirming that the stopped-flow method used was appropriate for the reaction. Single turnover stopped-flow experiments were used to identify catalytic intermediates. These data were globally fit confirming a three-step kinetic model. Independent absorption spectra acquired between 0.005 and 0.5 s of the reaction reveal a significant increase in absorbance at 375, 460, and 550 nm along with the hypsochromic shift of an Fe3+←S ligand-to-metal charge transfer band from 700 to 650 nm. The observed UV-visible absorption bands for the Fe3+-nitrile intermediate species are similar to low spin Fe3+-enzyme and model complexes bound by NO or N3. These data provide spectroscopic evidence for the direct coordination of the nitrile substrate to the nitrile hydratase active site low spin Fe3+ center.  相似文献   

14.
We studied anionic inhibition of the reaction CO2 + OH?? HCO3? catalyzed by human red cell carbonic anhydrase B (I) and C (II), using iodide and cyanate. In the forward reaction with respect to CO2 as the substrate, inhibition was mixed but favoring noncompetitive; the back reaction, with HCO3? as the substrate, yielded strict competitive kinetics. Mean inhibition constants, KI, in the pH range 7.2–7.5 are: iodide, 0.5 mm for enzyme B and 16 mm for C; cyanate, 0.8 μm for B and 20 μm for C. When OH? was considered as the substrate for the forward reaction, cyanate and chloride behaved as competitive inhibitors. The true inhibition constant (KI0) for cyanate (calculated for infinitely low OH?) is 0.4 μm for enzyme B and 4 μm for C. Apart from the difference in anion affinity and some 10-fold higher activity of C > B, the isozymes showed similar patterns of inhibition. Data agree with generally proposed mechanisms describing the active site as ZnH2O with pKa of about 7.  相似文献   

15.
l-Glycerol 3-phosphate dehydrogenase has been isolated and partially purified from the endosperm of developing castor beans. The enzyme is entirely cytosolic and is not found in the plastid fraction. No activity was found in germinating castor beans. The pH optimum for the reduction of dihydroxyacetone phosphate is 8.1 and is 9.6 for the reverse reaction. The molecular weight determined by gel filtration chromatography is between 71,000 and 83,000. Both substrates show substrate inhibition at concentrations about 13 μm for NADH and 400 μm for dihydroxyacetone phosphate. Substrate interaction kinetics gave limiting Km values of 2.7 and 35.5 μm for NADH and dihydroxyacetone phosphate, respectively. Substrate interaction and product inhibition kinetics were consistent with an ordered sequential mechanism with NADH being the first substrate to bind and NAD+ being the last product to dissociate.  相似文献   

16.
Thiazole synthase in Escherichia coli is an αβ heterodimer of ThiG and ThiH. ThiH is a tyrosine lyase that cleaves the Cα–Cβ bond of tyrosine, generating p-cresol as a by-product, to form dehydroglycine. This reactive intermediate acts as one of three substrates for the thiazole cyclization reaction catalyzed by ThiG. ThiH is a radical S-adenosylmethionine (AdoMet) enzyme that utilizes a [4Fe-4S]+ cluster to reductively cleave AdoMet, forming methionine and a 5′-deoxyadenosyl radical. Analysis of the time-dependent formation of the reaction products 5′-deoxyadenosine (DOA) and p-cresol has demonstrated catalytic behavior of the tyrosine lyase. The kinetics of product formation showed a pre-steady state burst phase, and the involvement of DOA in product inhibition was identified by the addition of 5′-methylthioadenosine/S-adenosylhomocysteine nucleosidase to activity assays. This hydrolyzed the DOA and changed the rate-determining step but, in addition, substantially increased the uncoupled turnover of AdoMet. Addition of glyoxylate and ammonium inhibited the tyrosine cleavage reaction, but the reductive cleavage of AdoMet continued in an uncoupled manner. Tyrosine analogues were incubated with ThiGH, which showed a strong preference for phenolic substrates. 4-Hydroxyphenylpropionic acid analogues allowed uncoupled AdoMet cleavage but did not result in further reaction (Cα–Cβ bond cleavage). The results of the substrate analogue studies and the product inhibition can be explained by a mechanistic hypothesis involving two reaction pathways, a product-forming pathway and a futile cycle.  相似文献   

17.
A partially-purified diacylglycerol (DG) lipase from bovine aorta has been characterized with respect to the effects of lipid metabolites and two lipase inhibitors, phenylboronic acid and tetrahydrolipstatin (THL). DG lipase activity was determined by the hydrolysis of the sn-1 position of 1-[1-4C]palmitoyl-2-oleoyl-sn-glycerol. The products of the lipase reaction, 2-monoacylglycerol (2-monoolein) and non-esterified fatty acids (oleate, arachidonate) produced a concentration-dependent (20–200 μM) inhibition of DG lipase activity. Oleoyl-CoA and dioleoylphosphatidic acid also inhibited aortic DG lipase activity, but lysophosphatidylcholine had little or no effect. The inhibition of aortic DG lipase by phenylboronic acid was competitive, with a Ki of approx. 4 mM. THL was a very potent inhibitor of aortic DG lipase; the concentration required for inhibition to 50% of control was 2–6 nM. THL was a very potent inhibitor of concentration of substrate in the assay was increased. Attempts to identify the aortic DG lipase by covalent-labelling with [14C]THL were unsuccessful. Immunoblotting experiments revealed that hormone-sensitive triacylglycerol lipase (HSL) could not be detected in bovine aorta.  相似文献   

18.
The reaction of glutamate dehydrogenase (l-glutamate: NAD+ oxidoreductase (deaminating) EC 1.4.1.2) from lupin nodules has been investigated in the direction of deamination by means of steady state velocity studies in the absence of products and inhibition studies with products and substrate analogs. The results are qualitatively and quantitatively consistent with a fully ordered reaction mechanism in which NAD+ binds to the enzyme first followed by l-glutamate. The order of product release is proposed to be NH4+ followed by 2-oxoglutarate and then NADH. In addition, product inhibition data indicate the formation of an enzyme-NAD-oxoglutarate dead-end complex.  相似文献   

19.
Differences between CTP and ATP as substrates for the (Na + K)-ATPase   总被引:1,自引:0,他引:1  
CTP was a poorer substrate than ATP when substituted in the (Na + K)-ATPase reaction assay, not only in terms of Km but also of V. CDP was a poorer inhibitor than ADP, so product inhibition cannot account for CTP being a poorer substrate. In the Na-ATPase reaction, which the enzyme also catalyzes, substituting CTP for ATP resulted in greater activity, arguing against CTP being less effective than ATP in forming the enzyme-phosphate intermediate common to both reactions. Ligands that favor the E2 conformational state of the enzyme, K+, Mg2+, and Mn2+, inhibited the (Na + K)-CTPase reaction more than the (Na + K)-ATPase. Conversely, Triton X-100, which favors the E1 conformational state of the enzyme, K+, Mg2+, and Mn2+, inhibited the (Na + K)-CTPase ATPase reaction but stimulated the (Na + K)-CTPase. Although the (Na + K)-ATPase reaction sequence probably involves cyclical interconversion between E1 and E2 conformational states (and is thus inhibitable by ligands favoring either state), the K-phosphatase reaction catalyzed by the enzyme apparently functions entirely in the E2 state. This reaction is better stimulated by CTP plus Na+ than by ATP plus Na+; moreover, CTP lessens inhibition by Triton X-100, and ATP lessens inhibition by inorganic phosphate (which reacts with the E2 state). These observations indicate that CTP is a poorer substrate than ATP because it is less effective in promoting conversion of E2 to E1, essential for the (Na + K)-dependent reaction mechanism. However, contrary to this rationale, dimethyl sulfoxide stimulated the (Na + K)-CTPase reaction although by other criteria, including inhibition of the (Na + K)-ATPase, the reagent appears to favor the E2 over the E1 conformational state.  相似文献   

20.
Initial velocity steady-state substrate kinetics for ATP phosphoribosyltransferase were determined in the direction reverse to the biosynthetic reaction and are consistent with a sequential kinetic mechanism. Histidine inhibited the reverse reaction cooperatively and completely. Product and alternate product inhibition studies were conducted to elucidate binding order. The alternate product β,γ-methylene ATP was competitive with respect to N1-phosphoribosyl-ATP and noncompetitive with respect to pyrophosphate. Phosphoribosylpyrophosphate was noncompetitive with respect to both substrates. These data and those of the biosynthetic direction reaction are in satisfactory quantitative agreement with the ordered Bi-Bi kinetic mechanism with ATP or phosphoribosyl-ATP binding to free 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