首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Incubation of whole cells of the nitrifying bacterium Nitrosomonas europaea with ethylene led to the formation of ethylene oxide. Ethylene oxide production was prevented by inhibitors of ammonium ion oxidation, and showed properties implying that ethylene is a substrate for the ammonia oxidising enzyme, ammonia monooxygenase. Endogenous substrates, hydroxylamine, hydrazine and ammonium ions were compared as sources of reducing power in terms of rates and stoichiometries of ethylene oxidation. The highest rates of ethylene oxide formation (15 mol h-1 mg protein-1) were obtained with hydrazine as donor. The data suggest that at high concentrations of ethylene the rate of oxidation is limited by the rate at which reducing power can be supplied to the monooxygenase, not by an intrinsic V max. Ethylene had an inhibitory effect on the rate of ammonium ion utilisation; an approximate K i of 80 M was derived, but the results deviated from simple competitive behaviour. Measurement of relative rates of ethylene oxide formation and ammonium ion utilization led to a k cat/K m value for ethylene of 1.1 relative to NH 4 + , or 0.04 relative to the true natural substrate, NH3. The effects of higher concentrations of ethylene oxide on oxygen uptake rates were also investigated. The results imply that ethylene oxide is also a substrate for the monooxygenase, but with a much lower affinity than ethylene.  相似文献   

2.
Kinetic studies on ammonia and methane oxidation by Nitrosococcus oceanus   总被引:2,自引:0,他引:2  
The kinetics of ammonia oxidation and the ability of a marine ammonia-oxidizing bacterium, Nitrosococcus oceanus, to metabolize methane were investigated in semicontinuous batch culture. The effects of inhibitors (acetylene and nitrapyrin) and coreactants were determined in order to elucidate the behavior of the ammonia oxygenase enzyme in N. oceanus. Acetylene and nitrapyrin were potent inhibitors and their effects were not mitigated by increased ammonia concentrations. Oxygen concentration had the effect of a mixed-type inhibitor; reduced oxygen inhibited the rate or ammonia oxidation at high substrate concentration but may enhance the rate at low substrate concentrations. Substrate affinity in terms of NH 4 + increased (K m decreased) with increasing pH. Optimal pH was about 8. Methane inhibited ammonia oxidation; the interaction was not simple competitive inhibition and the presence of multiple active sites on the enzyme was indicated by the behaviour of the inhibited treatments. Half-saturation constants for methane (K i=6.6 M) and ammonia (K m=8.1 M) were similar. N. oceanus oxidized methanol and methane linearly over time, with CO2 and cell material being produced at approximately equal rates.  相似文献   

3.
Desulfobacter postgatei is an acetate-oxidizing, sulfate-reducing bacterium that metabolizes acetate via the citric acid cycle. The organism has been reported to contain a si-citrate synthase (EC 4.1.3.7) which is activated by AMP and inorganic phosphate. It is show now, that the enzyme mediating citrate formation is an ATP-citrate lyase (EC 4.1.3.8) rather than a citrate synthase. Cell extracts (160,000xg supernatant) catalyzed the conversion of oxaloacetate (apparent K m=0.2 mM), acetyl-CoA (app. K m=0.1 mM), ADP (app. K m=0.06 mM) and phosphate (app. K m=0.7 mM) to citrate, CoA and ATP with a specific activity of 0.3 mol·min-1·mg-1 protein. Per mol citrate formed 1 mol of ATP was generated. Cleavage of citrate (app. K m=0.05 mM; V max=1.2 mol · min-1 · mg-1 protein) was dependent on ATP (app. K m=0.4 mM) and CoA (app. K m=0.05 mM) and yielded oxaloacetate, acetyl-CoA, ADP, and phosphate as products in a stoichiometry of citrate:CoA:oxaloacetate:ADP=1:1:1:1. The use of an ATP-citrate lyase in the citric acid cycle enables D. postgatei to couple the oxidation of acetate to 2 CO2 with the net synthesis of ATP via substrate level phosphorylation.  相似文献   

4.
Improved rates of ethylene oxidation by cell-free preparations from cotyledons of Vicia faba L. have been obtained using cryogenic storage techniques and by developing a method for the hydrolysis of ethylene oxide. Gel permeation chromatography showed that a low-molecular-size fraction was required for activity; accordingly, the kinetics of ethylene oxidation in the presence of this fraction were studied. Reduced pyridine nucleotides could substitute for the low-molecular-size fraction. Activity under a nitrogen atmosphere was 60% lower than in air. The need for reduced nicotinamide adenine dinucleotide phosphate (NADPH) and oxygen indicated that the enzyme might be a mixed-function oxidase. Using sufficient NADPH to approach saturation, the apparent Michaelis constant (K m) for ethylene was 1.94±0.38 · 10-8 M (aqueous phase), and when ethylene was saturating, the K m for NADPH was 3.7 · 10-5 M. Carbon monoxide was found to inhibit by competing with ethylene, and the inhibitor constant was 5.97 · 10-7 M in solution. In the presence of excess ethylene and NADPH, activity was highest in phosphate-buffered medium pH 7.9. The bulk of the activity was found in a microsomal fraction.Abbreviations Epps N-2-hydroxyethylpiperazine-N-3-propane sulphinic acid - Tris 2-amino-2-(hydroxymethyl)-1,3-porpanediol  相似文献   

5.
The peroxidase-catalyzed oxidation of 3,3",5,5"-tetramethylbenzidine (TMB), ortho-phenylenediamine (PDA), and 5-aminosalicylic acid (5-ASA) is significantly accelerated in the presence of 2-aminothiazole (AT) and melamine (MA), and an increase in their concentrations is associated with a parallel increase in the k cat and K m values for TMB and PDA. The activation of the peroxidase-catalyzed oxidation of TMB and PDA is quantitatively characterized by a coefficient (degree) (M–1) which significantly depends on pH in the range 6.2-6.4, 6.4-7.4, and 6.0-7.4 for the TMB–AT, TMB–MA, and PDA–MA pairs, respectively. An increase in the coefficient with increase in pH confirms nucleophilicity of activation of the peroxidase-catalyzed oxidation of the aromatic amines in the presence of AT and MA. Under optimal conditions the coefficients for the TMB–AT, PDA–AT, TMB–MA, and PDA–MA pairs vary in the limits of (1.90-3.53)·103 M–1.  相似文献   

6.
A comparative study of secondary specificities of enteropeptidase and trypsin was performed using peptide substrates with general formula A-(Asp/Glu) n -Lys(Arg)--B, where n = 1-4. This was the first study to demonstrate that, similar to other serine proteases, enteropeptidase has an extended secondary binding site interacting with 6-7 amino acid residues surrounding the peptide bond to be hydrolyzed. However, in the case of typical enteropeptidase substrates containing four negatively charged Asp/Glu residues at positions P2-P5, electrostatic interaction between these residues and the secondary site Lys99 of the enteropeptidase light chain is the main factor that determines hydrolysis efficiency. The secondary specificity of enteropeptidase differs from the secondary specificity of trypsin. The chromophoric synthetic enteropeptidase substrate G5DK-F(NO2)G (k cat/K m = 2380 mM–1·min–1) is more efficient than the fusion protein PrAD4K-P26 (k cat/K m = 1260 mM–1·min–1).  相似文献   

7.
Summary The apparent Michaelis constant (K m) of NADH for muscle-type (M4 isozyme) lactate dehydrogenases (LDHs) is highest, at any given temperature of measurement, for LDHs of cold-adapted vertebrates (Table 1). However, these interspecific differences in theK m of NADH are not due to variations in LDH-NADH binding affinity. Rather, theK m differences result entirely from interspecific variation in the substrate turnover constant (k cat) (Fig. 1; Table 2). This follows from the fact that theK m of NADH is equal tok cat divided by the on constant for NADH binding to LDH,k 1, so that interspecific differences ink cat, combined with identical values fork 1 among different LDH reactions, make the magnitude of theK m of NADH a function of substrate turnover number. The temperature dependence of theK m of NADH for a single LDH homologue is the net result of temperature dependence of bothk cat andk 1 (Figs. 3 and 4). Temperature independentK m values can result from simultaneous, and algebraically offsetting, increases ink cat andk 1 with rising temperature. Salt-induced changes in theK m of NADH also may be due to simultaneous perturbation of bothk cat andk 1 (Table 3). These findings are discussed from the standpoint of the evolution of LDH kinetic properties, particularly the interspecific conservation of catalytic and regulatory functions, in differently-adapted species.  相似文献   

8.
Two extracellular tannin acyl hydrolases (TAH I and TAH II) produced by an Antarctic filamentous fungus Verticillium sp. P9 were purified to homogeneity (7.9- and 10.5-fold with a yield of 1.6 and 0.9%, respectively) and characterized. TAH I and TAH II are multimeric (each consisting of approximately 40 and 46 kDa sub-units) glycoproteins containing 11 and 26% carbohydrates, respectively, and their molecular mass is approximately 155 kDa. TAH I and TAH II are optimally active at pH of 5.5 and 25 and 20°C, respectively. Both the enzymes were activated by Mg2+and Br ions and 0.5–2.0 M urea and inhibited by other metal ions (Zn2+, Cu2+, K+, Cd2+, Ag+, Fe3+, Mn2+, Co2+, Hg2+, Pb2+ and Sn2+), anions, Tween 20, Tween 60, Tween 80, Triton X-100, sodium dodecyl sulphate, β-mercaptoethanol, α-glutathione and 4-chloromercuribenzoate. Both tannases more efficiently hydrolyzed tannic acid than methyl gallate. E a of these reactions and temperature dependence (at 0–30°C) of k cat, k cat/K m, ΔG*, ΔH* and ΔS* for both the enzymes and substrates were determined. The k cat and k cat/K m values (for both the substrates) were considerably higher for the combined preparation of TAH I and TAH II.  相似文献   

9.
D-Lactate dehydrogenase (D-LDH) from Pediococcus pentosaceus ATCC 25745 was found to produce D-3-phenyllactic acid from phenylpyruvate. The optimum pH and temperature for enzyme activity were pH 5.5 and 45 °C. The Michaelis-Menten constant (K m), turnover number (k cat), and catalytic efficiency (k cat?K m) values for the substrate phenylpyruvate were estimated to be 1.73 mmol/L, 173 s?1, and 100 (mmol/L)?1 s?1 respectively.  相似文献   

10.
We identified and characterized a malate dehydrogenase from Streptomyces coelicolor A3(2) (ScMDH). The molecular mass of ScMDH was 73,353.5 Da with two 36,675.0 Da subunits as analyzed by matrix-assisted laser-desorption ionization–time-of-flight mass spectrometry (MALDI-TOF-MS). The detailed kinetic parameters of recombinant ScMDH are reported here. Heat inactivation studies showed that ScMDH was more thermostable than most MDHs from other organisms, except for a few extremely thermophile bacteria. Recombinant ScMDH was highly NAD+-specific and displayed about 400-fold (k cat) and 1,050-fold (k cat?K m) preferences for oxaloacetate reduction over malate oxidation. Substrate inhibition studies showed that ScMDH activity was inhibited by excess oxaloacetate (K i=5.8 mM) and excess L-malate (K i=12.8 mM). Moreover, ScMDH activity was not affected by most metal ions, but was strongly inhibited by Fe2+ and Zn2+. Taken together, our findings indicate that ScMDH is significantly thermostable and presents a remarkably high catalytic efficiency for malate synthesis.  相似文献   

11.
Bacterial chitosanases share weak amino acid sequence similarities at certain regions of each enzyme. These regions have been assumed to be important for catalytic activities of the enzyme. To verify this assumption, the functional importance of the conserved region in a novel thermostable chitosanase (TCH-2) from Bacillus coagulans CK108 was investigated. Each of the conserved amino acid residues (Leu64, Glu80, Glu94, Asp98, and Gly108) was changed to aspartate and glutamine or asparagine and glutamate by site-directed mutagenesis, respectively. Kinetic parameters for colloidal chitosan hydrolysis were determined with wild-type and 10 mutant chitosanases. The Leu64 Arg and Leu64 Gln mutations were essentially inactive and kinetic parameters such as V max and k cat were approximately 1/107 of those of the wild-type enzyme. The Asp98 Asn mutation did not affect the K m value significantly, but decreased k cat to 15% of that of wild-type chitosanase. On the other hand, the Asp98 srarr; Glu mutation affected neither K m nor k cat. The observation that approximately 15% of activity remained after the substitution of Asp98 by Asn indicated that the carboxyl side chain of Asp98 is not absolutely required for catalytic activity. These results indicate that the Leu64 residue is directly involved in the catalytic activity of TCH-2.  相似文献   

12.
Aldehyde dehydrogenase ST0064, the closest paralog of previously characterized allosteric non-phosphorylating glyceraldehyde-3-phosphate (GAP) dehydrogenase (GAPN, ST2477) from a thermoacidophilic archaeon, Sulfolobus tokodaii, was expressed heterologously and characterized in detail. ST0064 showed remarkable activity toward succinate semialdehyde (SSA) (K m of 0.0029 mM and k cat of 30.0 s?1) with no allosteric regulation. Activity toward GAP was lower (K m of 4.6 mM and k cat of 4.77 s?1), and previously predicted succinyl-CoA reductase activity was not detected, suggesting that the enzyme functions practically as succinate semialdehyde dehydrogenase (SSADH). Phylogenetic analysis indicated that archaeal SSADHs and GAPNs are closely related within the aldehyde dehydrogenase superfamily, suggesting that they are of the same origin.  相似文献   

13.
Peroxidase-catalyzed oxidation of o-phenylenediamine (PDA) is greatly activated with melamine (MA) in 15 mM phosphate–citrate buffer at pH 6.0–7.4 in a noncompetitive manner: k cat and K m increase in direct proportion to the MA concentration. An extent of the activation is quantitatively characterized with a coefficient (in M–1), which essentially increases along with the rise in pH from 6.0 to 7.4. MA acts as a nucleophilic catalyst in the oxidation process: it most likely affects the peroxidase active site from the distal position of heme. MA noncompetitively inhibits the peroxidase oxidation of PDA at pH 4.3, since it completely loses its nucleophilic properties in acidic medium. A rapid, highly accurate, and simple analytical test system based on the kinetics of melamine-activated oxidation of PDA is proposed for the quantitative determination of melamine within the concentration range of 10–4–10–3 M. This test system uses the spectrophotometric determination of the PDA oxidation product at 455 nm.  相似文献   

14.
Intact cells ofNitrosomonas europaea grown in an ammonium salts medium will oxidise ammonium ions, hydroxylamine and ascorbate-TMPD; there is no oxidation of carbon monoxide, methane or methanol. TheK m value for ammonia oxidation is highly pH dependent with a minimum value of 0.5 mM above pH 8.0. This suggests that free ammonia is the species crossing the cytoplasmic membrane(s). The measurement of respiration driven proton translocation indicates that there is probably only one proton translocating loop (loop 3) association with hydroxylamine oxidation. The oxidation of endogenous substrates is sometimes associated with more than one proton-translocating loop. These results indicate that during growth hydroxylamine oxidation is probably associated with a maximum P/O ratio of 1.Abbreviations H+/O ratio g equiv. H+ translocated/g atom O consumed  相似文献   

15.
Soluble extracts of Methylococcus capsulatus (Bath) that readily oxidise methane to methanol will also oxidise ammonia to nitrite via hydroxylamine. The ammonia oxidising activity requires O2, NADH and is readily inhibited by methane and specific inhibitors of methane mono-oxygenase activity. Hydroxylamine is oxidised to nitrite via an enzyme system that uses phenazine methosulphate (PMS) as an electron acceptor. The estimated K mvalue for the ammonia hydroxylase activity was 87 mM but the kinetics of the oxidation were complex and may involve negative cooperativity.Abbreviations PMS Phenazine methosulphate - NADH nicotinamide adenine dinucleotide, reduced form - K m Michaelis constant - NO 2 - nitrite - NH2OH hydroxylamine  相似文献   

16.
We previously isolated from Nicotiana glutinosa leaves three distinct cDNA clones, NGR1, NGR2, and NGR3, encoding a wound-inducible RNase NW, and putative RNases NGR2 and NGR3, respectively. In this study, we produced RNases NW and NGR3 in Escherichia coli and purified them to homogeneity. RNase NGR3 had non-absolute specificity toward polynucleotides, although RNase NW preferentially cleaved polyinosinic acid (Poly I). Both RNases NW and NGR3 were more active toward diribonucleoside monophosphates ApG, CpU, and GpU. Furthermore, kinetic parameters for RNase NW (K m, 0.778 mM and k cat, 1938 min?1) and RNase NGR3 (K m, 0.548 mM and k cat, 408 min?1) were calculated using GpU as a substrate.  相似文献   

17.
Acinetobacter strain IVS-B aerobically grows on isovalerate as sole carbon and energy source. Isovalerate is metabolised via isovaleryl-CoA, an intermediate of the oxidative (S)-leucine degradation pathway. A 3-methylglutaconyl-CoA hydratase (EC 4.2.1.18) was purified 65-fold to apparent homogeneity from cell-free extracts of isovalerate-grown cells of Acinetobacter strain IVS-B. The enzyme was found to be a homotetramer (115.2 kDa) composed of four identical subunits of 28.8 kDa not containing any cofactors. The enzyme was shown to catalyse the hydration of (E)-glutaconyl-CoA (k cat=18 s−1, K m=40 μM) and the dehydration of (S)-3-hydroxyglutaryl-CoA (k cat=13 s−1, K m=52 μM), albeit with somewhat lower catalytic efficiencies as compared to the 3-methyl derivatives, 3-methylglutaconyl-CoA (k cat=138 s−1, K m=14 μM) and (S)-3-hydroxy-3-methylglutaryl-CoA (k cat=60 s−1, K m=36 μM). Thus, the mechanistically simple syn-addition of water to the (E)-isomer of 3-methylglutaconyl-CoA of the leucine degradative pathway leading to the common intermediate (S)-3-hydroxy-3-methylglutaryl-CoA was assigned as the major physiological role to this enzyme. The amino acid sequence of 3-methylglutaconyl-CoA hydratase from Acinetobacter sp. was found to be related to over 100 prokaryotic enoyl-CoA hydratases (up to 50% identity), possibly all being 3-methylglutaconyl-CoA hydratases.An erratum to this article can be found at  相似文献   

18.
The catalytic activity of the microperoxidase-8/H2O2 system toward tyramine and 3-(4-hydroxyphenyl)propionic acid has been determined in acetate buffer, pH 5.0. Operating with a strong excess of hydrogen peroxide, the rate-determining step of the reaction was substrate oxidation. Owing to the fast microperoxidase-8 degradation, only the very initial phase of the reactions were analyzed. The reaction rates follow a substrate saturation behavior, with turnover numbers [kcat=26±1 s–1 for 3-(4-hydroxyphenyl)propionic acid and kcat=22±1 s–1 for tyramine] that were similar for the two substrates. In contrast, the KM values indicated a reduced affinity for the catalyst active species by the positively charged phenol, probably due to repulsive interaction with the protonated N-terminal microperoxidase-8 amino group. The reactivity of the catalyst active species was studied upon incubation of microperoxidase-8 with a small excess hydrogen peroxide, followed by reaction with the phenolic substrates. The kinetic analysis showed that more than two active species are accumulated. The species responsible for the faster reactions was present in solution as a minor fraction. The active intermediate which accumulated in a larger amount (intermediate III) has a reduced substrate oxidation activity. Comparison of this activity with the kinetic constants obtained under turnover experiments shows that intermediate III is not involved in the microperoxidase-8 catalytic cycle. The active species of the catalytic process are intermediates I and II, which in the absence of substrate rapidly convert to intermediate III.Abbreviations ABTS 2,2-azinobis(3-ethylbenzothiazoline-6-sulfonic acid) - HPA 3-(4-hydroxyphenyl)propionic acid - HRP horseradish peroxidase - MP-8 microperoxidase-8  相似文献   

19.
An aspartic protease that is significantly produced by baculovirus-infected Spodoptera frugiperda Sf9 insect cells was purified to homogeneity from a growth medium. To monitor aspartic protease activity, an internally quenched fluoresce (IQF) substrate specific to cathepsin D was used. The purified aspartic protease showed a single protein band on SDS–PAGE with an apparent molecular mass of 40 kDa. The N-terminal amino acid sequence of the enzyme had a high homology to a Bombyx mori aspartic protease. The enzyme showed greatest affinity for the IQF substrate at pH 3.0 with a K m of 0.85 μM. The k cat and k cat?K m values were 13 s?1 and 15 s?1 μM?1 respectively. Pepstatin A proved to be a potent competitive inhibitor with inhibitor constant, K i, of 25 pM.  相似文献   

20.
Crude bromelain extracts from pineapple stems (Ananas comosus) were fractionated by two-step FPLC-cation-exchange chromatography. At least eight basic proteolytically active components were detected. The two main components F4 and F5 together with the most active proteinase fraction F9 were characterized by SDS-PAGE, mass spectroscopy, multizonal cathodal electrophoresis, partial amino acid sequence, and monosaccharide composition analysis. F9 amounts to about 2% of the total protein and has a 15 times higher specific activity against the substratel-pyroglutamyl-l-phenylanalyl-l-leucine-p-nitroanilide (PFLNA) than the main component F4. The molecular masses of F4, F5, and F9 were determined to 24,397, 24,472, and 23,427, respectively, by mass spectroscopy. Partial N-terminal amino acid sequence analysis (20 amino acids) revealed that F9 differs from the determined sequence of F4 and F5 by an exchange at position 10 (tyrosineserine) and position 20 (asparagine glycine). F4 and F5 contained fucose, N-acetylglucosamine, xylose, and mannose in ratio of 1.02.01.02.0, but only 50% of the proteins seem to be glycosylated, whereas F9 was found to be unglycosylated. Polyclonal antibodies (IgG) against F9 detected F4 and F5 with tenfold reduced reactivity. ThepH optimum of F4 and F5 was betweenpH4.0 and 4.5 and for F9 close to neutralpH. The kinetic parameters for PFLNA hydrolysis were similar for F4 (K m 2.30 mM,k cat 0.87 sec–1 and F5 (K m 2.42 mM,k cat 0.68 sec–1), and differed greatly from F9 (K m 0.40 mM,k cat 3.94 sec–1).Dedicated to H. Tschesche, Bielefeld, Germany, on behalf of his 60th anniversary.  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