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.
C4a-hydroperoxyflavin is found commonly in the reactions of flavin-dependent monooxygenases, in which it plays a key role as an intermediate that incorporates an oxygen atom into substrates. Only recently has evidence for its involvement in the reactions of flavoprotein oxidases been reported. Previous studies of pyranose 2-oxidase (P2O), an enzyme catalyzing the oxidation of pyranoses using oxygen as an electron acceptor to generate oxidized sugars and hydrogen peroxide (H(2)O(2)), have shown that C4a-hydroperoxyflavin forms in P2O reactions before it eliminates H(2)O(2) as a product (Sucharitakul, J., Prongjit, M., Haltrich, D., and Chaiyen, P. (2008) Biochemistry 47, 8485-8490). In this report, the solvent kinetic isotope effects (SKIE) on the reaction of reduced P2O with oxygen were investigated using transient kinetics. Our results showed that D(2)O has a negligible effect on the formation of C4a-hydroperoxyflavin. The ensuing step of H(2)O(2) elimination from C4a-hydroperoxyflavin was shown to be modulated by an SKIE of 2.8 ± 0.2, and a proton inventory analysis of this step indicates a linear plot. These data suggest that a single-proton transfer process causes SKIE at the H(2)O(2) elimination step. Double and single mixing stopped-flow experiments performed in H(2)O buffer revealed that reduced flavin specifically labeled with deuterium at the flavin N5 position generated kinetic isotope effects similar to those found with experiments performed with the enzyme pre-equilibrated in D(2)O buffer. This suggests that the proton at the flavin N5 position is responsible for the SKIE and is the proton-in-flight that is transferred during the transition state. The mechanism of H(2)O(2) elimination from C4a-hydroperoxyflavin is consistent with a single proton transfer from the flavin N5 to the peroxide leaving group, possibly via the formation of an intramolecular hydrogen bridge.  相似文献   

2.
Dinitrogenase from a nifV mutant of Klebsiella pneumoniae contains an altered form of iron-molybdenum cofactor (FeMoco) that lacks a biologically active homocitric acid molecule. Change in the composition of FeMoco led to substantial variation in the kinetics of nitrogenase action. The KmS of the mutant enzyme for N2 and N2O were 0.244 and 0.175 atm (24,714 and 17,726 kPa), respectively. The km for N2 was higher and the Km for N2O was lower than that for the wild-type enzyme. The mutant enzyme was ineffective in N2 fixation, in N2O reduction, and in HD formation, as indicated by the low Vmax of these reactions with saturating levels of substrate and under conditions of saturating electron flux. These observations provide further support for the concept that N2, N2O, and D2 interact with the same form of dinitrogenase. H2 evolution by the mutant enzyme is only partially inhibited by CO. Observation that different numbers of electrons are stored in CO-inhibited than in noninhibited dinitrogenase before H2 is released suggests that the mutant enzyme has more sites responsible for H2 evolution than the wild-type enzyme, whose H2 evolution is not inhibited by CO.  相似文献   

3.
Dissimilatory reduction of NO(2) to N(2)O and NH(4) by a soil Citrobacter sp. was studied in an attempt to elucidate the physiological and ecological significance of N(2)O production by this mechanism. In batch cultures with defined media, NO(2) reduction to NH(4) was favored by high glucose and low NO(3) concentrations. Nitrous oxide production was greatest at high glucose and intermediate NO(3) concentrations. With succinate as the energy source, little or no NO(2) was reduced to NH(4) but N(2)O was produced. Resting cell suspensions reduced NO(2) simultaneously to N(2)O and free extracellular NH(4). Chloramphenicol prevented the induction of N(2)O-producing activity. The K(m) for NO(2) reduction to N(2)O was estimated to be 0.9 mM NO(2), yet the apparent K(m) for overall NO(2) reduction was considerably lower, no greater than 0.04 mM NO(2). Activities for N(2)O and NH(4) production increased markedly after depletion of NO(3) from the media. Amendment with NO(3) inhibited N(2)O and NH(4) production by molybdate-grown cells but not by tungstate-grown cells. Sulfite inhibited production of NH(4) but not of N(2)O. In a related experiment, three Escherichia coli mutants lacking NADH-dependent nitrite reductase produced N(2)O at rates equal to the wild type. These observations suggest that N(2)O is produced enzymatically but not by the same enzyme system responsible for dissimilatory reduction of NO(2) to NH(4).  相似文献   

4.
Under anaerobic conditions, Klebsiella pneumoniae reduced nitrite (NO2-), yielding nitrous oxide (N2O) and ammonium ions (NH4+) as products. Nitrous oxide formation accounted for about 5% of the total NO2- reduced, and NH4+ production accounted for the remainder. Glucose and pyruvate were the electron donors for NO2- reduction to N2O by whole cells, whereas glucose, NADH, and NADPH were found to be the electron donors when cell extracts were used. On the one hand, formate failed to serve as an electron donor for NO2- reduction to N2O and NH4+, whereas on the other hand, formate was the best electron donor for nitrate reduction in either whole cells or cell extracts. Mutants that are defective in the reduction of NO2- to NH4+ were isolated, and these strains were found to produce N2O at rates comparable to that of the parent strain. These results suggest that the nitrite reductase producing N2O is distinct from that producing NH4+. Nitrous oxide production from nitric oxide (NO) occurred in all mutants tested, at rates comparable to that of the parent strain. This result suggests that NO reduction to N2O, which also uses NADH as the electron donor, is independent of the protein(s) catalyzing the reduction of NO2- to N2O.  相似文献   

5.
Nitrogen and oxygen transformations were studied in a bioturbated (reworked by animals) estuarine sediment (Norsminde Fjord, Denmark) by using a combination of N isotope (NO(3)), specific inhibitor (C(2)H(2)), and microsensor (N(2)O and O(2)) techniques in a continuous-flow core system. The estuarine water was NO(3) rich (125 to 600 muM), and NO(3) was consistently taken up by the sediment on the four occasions studied. Total NO(3) uptake (3.6 to 34.0 mmol of N m day) corresponded closely to N(2) production (denitrification) during the experimental steady state, which indicated that dissimilatory, as well as assimilatory, NO(3) reduction to NH(4) was insignificant. When C(2)H(2) was applied in the flow system, denitrification measured as N(2)O production was often less (58 to 100%) than the NO(3) uptake because of incomplete inhibition of N(2)O reduction. The NO(3) formed by nitrification and not immediately denitrified but released to the overlying water, uncoupled nitrification, was calculated both from NO(3) dilution and from changes in NO(3) uptake before and after C(2)H(2) addition. These two approaches gave similar results, with rates ranging between 0 and 8.1 mmol of N m day on the four occasions. Attempts to measure total nitrification activity by the difference between NH(4) fluxes before and after C(2)H(2) addition failed because of non-steady-state NH(4) fluxes. The vertical distribution of denitrification and oxygen consumption was studied by use of N(2)O and O(2) microelectrodes. The N(2)O profiles measured during the experimental steady state were often irregularly shaped, and the buildup of N(2)O after C(2)H(2) was added was much too fast to be described by a simple diffusion model. Only bioturbation by a dense population of infauna could explain these observations. This was corroborated by the relationship between diffusive and total fluxes, which showed that only 19 to 36 and 29 to 62% of the total O(2) uptake and denitrification, respectively, were due to diffusion-reaction processes at the regular sediment surface, excluding animal burrows.  相似文献   

6.
Durrant MC 《Biochemistry》2002,41(47):13946-13955
Quantum calculations have been used to examine the energetics of the reactions of diazene and isodiazene with H(2) and the properties of the Fe and Mo sites of the nitrogenase iron-molybdenum cofactor with respect to the binding of H and H(2). The results have been used to extend the model for N(2) reduction by nitrogenase given in the preceding paper to describe the formation of HD from D(2). The proposed mechanism for HD formation invokes a combination of two well-established chemical reactions, namely, competitive protonation of metal N(2) species at either the metal or at N(2), followed by scrambling of D(2) at a metal hydride. The model is evaluated against the available biochemical data for the nitrogenase HD formation reaction and extended to account for H(2) inhibition of N(2) reduction and the reduction of H(+) in the absence of other substrates.  相似文献   

7.
The reactions of cis-[PtCl(NH3)2(H2O)]+ with L-methionine have been studied by 1D 195Pt and 15N NMR, and by 2D[1H, 15N] NMR. When the platinum complex is in excess, the initial product, cis-[PtCl(NH3)2(Hmet-S)]+ undergoes slow ring closure to [Pt(NH3)2(Hmet-N,S)]2+. Slow ammine loss then occurs to give the isomer of [PtCl(NH3)(Hmet-N,S)]+ with chloride trans to sulfur. When methionine is in excess, a reaction sequence is proposed in which trans-[PtCl(NH3)(Hmet-S)2]+ isomerises to the cis-isomer, with subsequent ring closure reactions leading to cis-[Pt(Hmet-N,S)2]2+. Near pH 7, methionine is unreactive toward cis-[PtCl(OH)(NH3)2]. By contrast, L-cysteine reacts readily with cis-[PtCl(OH)(NH3)2] at pH 7, but there were many reaction products, including bridged species. Cis-[PtCl(OH)(NH3)2] reacts with reduced thiols in ultrafiltered plasma but these are oxidized if the plasma is not fresh or appropriately stored. With very low concentrations of the platinum complexes (35.5 microM), HPLC experiments (UV detection at 305 nm) indicate that the thiolate (probably cysteine) reactions become simpler as bridging becomes less important.  相似文献   

8.
Klebsiella pneumoniae nitrogenase reduced azide, at 30 degrees C and pH 6.8-8.2, to yield ammonia (NH3), dinitrogen (N2) and hydrazine (N2H4). Reduction of (15N = 14N = 14N)-followed by mass-spectrometric analysis showed that no new nitrogen-nitrogen bonds were formed. During azide reduction, added 15N2H4 did not contribute 15N to NH3, indicating lack of equilibration between enzyme-bound intermediates giving rise to N2H4 and N2H4 in solution. When azide reduction to N2H4 was partially inhibited by 15N2, label appeared in NH3 but not in N2H4. Product balances combined with the labelling data indicate that azide is reduced according to the following equations: (formula: see text); N2 was a competitive inhibitor and CO a non-competitive inhibitor of azide reduction to N2H4. The percentage of total electron flux used for H2 evolution concomitant with azide reduction fell from 26% at pH 6.8 to 0% at pH 8.2. Pre-steady-state kinetic data suggest that N2H4 is formed by the cleavage of the alpha-beta nitrogen-nitrogen bond to bound azide to leave a nitride (= N) intermediate that subsequently yields NH3.  相似文献   

9.
6-methylated guanine dinucleotides were used to study the influence of hydrogen bonding on the specific binding of the antitumor drug cDDP, cis-PtCl2(NH3)2, to DNA. In this interaction, the guanine-06 site appears to be important in explaining the preference for a pGpG-N7(1),N7(2) chelate, which results from H-bridge formation with the ammine ligand of cDDP. Guanine-06 methylated dinucleotides and the nonmodified dinucleotides were reacted with [Pt(dien)Cl]+, cis-PtCl2(NH3)2, and cis-[Pt(NH3)2(H2O)2]2+ and the reaction products were characterized by 1H NMR using pH titrations. Methylation at guanine-06 clearly reduces the preference for the guanine. In competition experiments monitored by NMR and experiments using UV spectrophotometry a decreasing reactivity towards [Pt(dien)(H2O)]2+ and cis-[Pt(NH3)2(H2O)2]2+ was found, in the order of d(GpG) greater than d(GomepG) greater than d(GpGome) greater than d(GomepGome). The difference in reactivity between 5' guanine methylation and 3' guanine methylation is ascribed to differences in the H-bond formation with the backbone phosphate. The resulting reduced stacking of the bases in both modified dinucleotides, compared to the bases in d(GpG), results in a preference for the 3' guanine over 5'.  相似文献   

10.
The O-antigen of Proteus mirabilis O38 was found to be unique among bacterial polysaccharides and to have the following structure: [carbohydrate structure in text] where D-Qui4N(Ac-D-Asp) is 4-(N-acetyl-D-aspart-4-ylamino)-4,6-dideoxy-D-glucose and AcEtnP is 2-acetamidoethyl phosphate. Neither of these entities have been hitherto found in natural polysaccharides. Structural studies were performed using 1D and 2D NMR spectroscopy, including experiments run in an H2O/D2O mixture to reveal correlations for NH protons. In addition, dephosphorylation, carboxyl reduction and selective cleavages were applied. Solvolysis of the polysaccharide with anhydrous HF gave an alpha-D-GlcNAc-(1-->3)-D-Qui4N(Ac-D-Asp) disaccharide. Solvolysis with trifluoromethanesulfonic (triflic) acid afforded D-GlcNAc6(AcEtnP), thus showing the suitability of this reagent for the preparation of phosphorylated sugar derivatives.  相似文献   

11.
To reduce nitrogenous emissions from composting, two different countermeasures were applied simultaneously in swine manure composting. One was forming struvite by adding Mg and P at the start of composting, and the other was to promote nitratation (nitrite being oxidized nitrate) by adding nitrite-oxidizing bacteria after the thermophilic phase of composting. In the laboratory- and mid-scale composting experiments, 25-43% of NH3, 52-80% of N2O and 96-99% of NO emissions were reduced. From the nitrogen balance, it was revealed that the struvite formation reduced not only NH3, but also other nitrogenous emissions except N2O. The amount of total nitrogen losses was reduced by 60% by the two combined countermeasures, against 51% by the struvite formation alone. However, the nitratation promotion dissolved struvite crystals due to the pH decline, diminishing the effect of struvite as a slow-release fertilizer.  相似文献   

12.
The effect of the partial pressure of oxygen (PO2) on the formation of vesicles, which are thought to be the site of N2 fixation in Frankia, was studied in HFPCcI3, an effective isolate from Casuarina cunninghamiana. Unlike other actinorhizal root nodules, vesicles are not produced by the endophyte in Casuarina nodules. However, in culture under aerobic conditions, large, phase-bright vesicles are formed in HFPCcI3 within 20 h following removal of NH+4 from the culture medium and reach peak numbers within 72 to 96 h. In vivo acetylene reduction activity parallels vesicle formation. Optimum rates of acetylene reduction in short-term assays occurred at 20% O2 (0.2 atm (1 atm = 101.325 kPa] in the gas phase. O2 uptake (respiration) determined polarographically showed diffusion-limited kinetics and remained unsaturated by O2 until 300 microM O2. In contrast, respiration in NH+4-grown cells was saturated by O2 between 8 and 10 microM O2. These results indicate the presence of a diffusion barrier associated with the vesicles. Vesicle development was repressed in cells incubated in N-free media sparged with gas mixtures with PO2 between 0.001 and 0.003 atm. Nitrogenase was induced under these conditions, but acetylene reduction was extremely O2 sensitive. The kinetics of O2 uptake as a function of dissolved O2 concentration in avesicular cells were similar to those in NH+4-grown cells indicating the lack of a diffusion barrier. These results demonstrate that vesicle formation and the development of the O2 protection mechanisms of nitrogenase are regulated by ambient PO2 in HFPCcI3.  相似文献   

13.
Abraham B  Hicks W  Jia Y  Baek JH  Miller JL  Alayash AI 《Biochemistry》2011,50(45):9752-9766
We have previously shown that hydrogen peroxide (H(2)O(2)) triggers irreversible oxidation of amino acids exclusive to the β-chains of purified human hemoglobin (HbAo). However, it is not clear, whether α- or β-subunit Hb variants exhibit different oxidative resistance to H(2)O(2) when compared to their native HbAo. Hb Providence contains two β-subunit variants with single amino acid mutations at βLys82→Asp (βK82D) and at βLys82→Asn (βK82N) positions and binds oxygen at lower affinity than wild type HbA. We have separated Hb Providence into its 3 component fractions, and contrasted oxidative reactions of its β-mutant fractions with HbAo. Relative to HbAo, both βK82N and βK82D fractions showed similar autoxidation kinetics and similar initial oxidation reaction rates with H(2)O(2). However, a more profound pattern of changes was seen in HbAo than in the two Providence fractions. The structural changes in HbAo include a collapse of β-subunits, and α-α dimer formation in the presence of excess H(2)O(2). Mass spectrometric and amino acid analysis revealed that βCys93 and βCys112 were oxidized in the HbAo fraction, consistent with oxidative pathways driven by a ferrylHb and its protein radical. These amino acids were oxidized at a lesser extent in βK82D fraction. While the 3 isolated components of Hb Providence exhibited similar ligand binding and oxidation reaction kinetics, the variant fractions were more effective in consuming H(2)O(2) and safely internalizing radicals through the ferric/ferryl pseudoperoxidase cycle.  相似文献   

14.
The means by which neutrophils within the body ward off infectious and neoplastic processes by the activation of molecular oxygen, as well as how such mechanisms dysfunction, is the subject of extensive ongoing research. Most previous studies of neutrophil activation indicate that there is a transient production of reactive oxygen species. Luminol-amplified chemiluminescence surveillance of O2-. and H2O2 supported these general findings. Yet, recent studies showed that production of reactive oxygen species by PMA-stimulated neutrophils is not transient but persistent; however, luminol-dependent methods do not corroborate such findings. The kinetics of O2-. production by human neutrophils were studied using luminol-amplified chemiluminescence (CL), spin trapping combined with electron spin resonance detection, and ferricytochrome c reduction. The effects of pH and O2 level on luminol-amplified CL were determined using hypoxanthine/xanthine oxidase to produce O2-. and H2O2 in cell-free systems. As we have found by electron spin resonance and ferricytochrome c reduction, stimulated neutrophils continued to generate O2-. for several hours, yet when luminol-amplified CL was used to continuously follow radical production, CL was shortly lost. Similar loss of CL was observed with continuous enzymatic formation of O2-. and H2O2. The failure of the CL assay to report O2-. and H2O2 formation results from some luminol reaction product which interferes with the light reaction. Our results show that the cells are operative for long periods indicating that cell exposure to prolonged O2-. fluxes does not terminate radical production, and even when pH, [O2], and reagents are optimized, the use of luminol-amplified CL is not a valid assay for continuous monitoring of O2-. and H2O2 generated by either stimulated neutrophils or in cell-free systems.  相似文献   

15.
The role of the glutamate dehydrogenase reaction as a pathway of glutamate synthesis was studied by incubating synaptosomes with 5 mM 15NH4Cl and then utilizing gas chromatography-mass spectrometry to measure isotopic enrichment in glutamate and aspartate. The rate of formation of [15N]glutamate and [15N]aspartate from 5 mM 15NH4Cl was approximately 0.2 nmol/min/mg of protein, a value much less than flux through glutaminase (4.8 nmol/min/mg of protein) but greater than flux through glutamine synthetase (0.045 nmol/min/mg of protein). Addition of 1 mM 2-oxoglutarate to the medium did not affect the rate of [15N]glutamate formation. O2 consumption and lactate formation were increased in the presence of 5 mM NH3, whereas the intrasynaptosomal concentrations of glutamate and aspartate were unaffected. Treatment of synaptosomes with veratridine stimulated reductive amination of 2-oxoglutarate during the early time points. The production of ([15N]glutamate + [15N]aspartate) was enhanced about twofold in the presence of 5 mM beta-(+/-)-2-aminobicyclo [2.2.1]heptane-2-carboxylic acid, a known effector of glutamate dehydrogenase. Supplementation of the incubation medium with a mixture of unlabelled amino acids at concentrations similar to those present in the extracellular fluid of the brain had little effect on the intrasynaptosomal [glutamate] and [aspartate]. However, the enrichment in these amino acids was consistently greater in the presence of supplementary amino acids, which appeared to stimulate modestly the reductive amination of 2-oxoglutarate. It is concluded: (a) compared with the phosphate-dependent glutaminase reaction, reductive amination is a relatively minor pathway of synaptosomal glutamate synthesis in both the basal state and during depolarization; (b) NH3 toxicity, at least in synaptosomes, is not referable to energy failure caused by a depletion of 2-oxoglutarate in the glutamate dehydrogenase reaction; and (c) transamination is not a major mechanism of glutamate nitrogen production in nerve endings.  相似文献   

16.
A comparative study of the binding of square planar cis- and trans-[Pt(NH3)2Cl2] complexes and the octahedral [Ru(NH3)5(H2O)]3+ complex to tRNAphe from yeast was carried out by X-ray crystallography. Both of the carcinostatic compounds, cis-[Pt(NH3)2Cl2] and [Ru(NH3)5(H2O)]3+ show similarities in their mode of binding to tRNA. These complexes bind specifically to the N(7) positions of guanines G15 and G18 in the dihydrouridine loop. [Ru(NH3)5(H2O)]3+ has an additional binding site at N(7) of residue G1 after extensive soaking times (58 days). A noncovalent binding site for ruthenium is also observed in the deep groove of the acceptor stem helix with shorter (25 days) soaking time. The major binding site for the inactive trans-[Pt(NH3)Cl2] complex is at the N(1) position of residue A73, with minor trans-Pt binding sites at the N(7) positions of residues Gm34, G18 and G43. The similarities in the binding modes of cis-[Pt(NH3)2Cl2] and [Ru(NH3)5(H2O)]3+ are expected to be related to their carcinostatic properties.  相似文献   

17.
We have isolated a protein factor from rat liver which stimulates anthranilamide hydroxylation by the microsomes in the presence of NADPH and oxygen and showed this factor to contain Cu and Zn and to have superoxide dismutase activity [Biochim. Biophys. Acta 365, 148-157 (1974)]. In the present study, this protein factor was confirmed to be a superoxide dismutase (SOD) by comparison of the recovery of SOD activity with that of anthranilamide hydroxylation-stimulating activity at each step of its purification, by inhibition of SOD activity with NaCN and hydrogen peroxide (H2O2), and by recovery of the SOD activity of the protein factor after reconstitution with Cu2+ and/or Zn2+. At a given SOD activity level, there was no difference among the rat liver SOD, Cu,Zn-SOD from bovine erythrocytes, and Mn-SOD from Serratia marcescens in their ability to stimulate anthranilamide hydroxylation not only by rat liver microsomes, but also by the reconstituted cytochrome P-450-containing monooxygenase system. Rat liver SOD stimulated anthranilamide hydroxylation by the reconstituted system in proportion to its amount below a protein concentration of 1 microgram/ml. In anthranilamide hydroxylation by the reconstituted system without SOD, only a slight hydroxylase activity was found at the initial stage of the reaction and a marked increase in the amounts of NADPH oxidized and H2O2 formed was observed after a lag time. In the presence of rat liver SOD, however, the hydroxylase activity was markedly and continuously increased almost proportionally to reaction time with a concomitant decrease in the amounts of NADPH oxidized and H2O2 formed. In addition, a trace of 3-OH anthranilamide, one of the products, not only stimulated NADPH-dependent H2O2 formation in the reconstituted system, but also inhibited the apparent reduction of cytochrome P-450 by NADPH in the reconstituted system. These effects of 3-OH anthranilamide were diminished by rat liver SOD. When a trace of 3-OH anthranilamide were added to a system composed of NADPH-cytochrome c (P-450) reductase and NADPH, H2O2 formation and NADPH oxidation were markedly stimulated. However, on addition of 3-OH anthranilamide to the system containing rat liver SOD, no stimulation on either H2O2 formation or NADPH oxidation was found.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
Prebiotic synthesis in atmospheres containing CH4, CO,and CO2   总被引:2,自引:0,他引:2  
The prebiotic synthesis of organic compounds using a spark discharge on various simulated primitive earth atmospheres at 25 degrees C has been studied. Methane mixtures contained H2 + CH4 + H2O + N2 + NH3 with H2/CH4 molar ratios from 0 to 4 and pNH3 = 0.1 torr. A similar set of experiments without added NH3 was performed. The yields of amino acids (1.2 to 4.7% based on the carbon) are approximately independent of the H2/CH4 ratio and whether NH3 was present, and a wide variety of amino acids are obtained. Mixtures of H2 + CO + H2O + N2 and H2 + CO2 + H2O + N2, with and without added NH3, all gave about 2% yields of amino acids at H2/CO and H2/CO2 ratios of 2 to 4. For a H2/CO2 ratio of 0, the yield of amino acids is extremely low (10(-3)%). Glycine is almost the only amino acid produced from CO and CO2 model atmospheres. These results show that the maximum yield is about the same for the three carbon sources at high H2/carbon ratios, but that CH4 is superior at low H2/carbon ratios. In addition, CH4 gives a much greater variety of amino acids than either CO or CO2. If it is assumed that an abundance of amino acids more complex than glycine was required for the origin of life, then these results indicate the requirement for CH4 in the primitive atmosphere.  相似文献   

19.
A second alternative nitrogenase complex (nitrogenase 3) was purified from a nifHDK deletion strain of Azotobacter vinelandii. The active complex is made up of two components, dinitrogenase 3 and dinitrogenase reductase 3. Dinitrogenase 3 contains two protein subunits (alpha, Mr 58,000, and beta, Mr 50,000) which assemble into at least two active configurations: alpha 2 beta 2 (dinitrogenase 3s) and alpha 1 beta 2 (dinitrogenase 3F). Dinitrogenase 3s contains 24 Fe and 18 acid-labile S2-ions per Mr 216,000, and dinitrogenase 3F contains 11 Fe and 9 acid-labile S2-ions per Mr 158,000. Dinitrogenase reductase 3 is composed of two protein subunits of identical Mr (32,500) and contains four Fe and four acid-labile S2- ions per Mr 65,000. On two-dimensional gels, the protein subunits of the nitrogenase 3 complex comigrated with the four Mo-, V-, and NH4+-repressible proteins originally designated as N2ase B: the nitrogenase hypothesized to exist in the alternative N2 fixation system first described in 1980 (P.E. Bishop, D. M. L. Jarlenski, and D. R. Hetherington, Proc. Natl. Acad. Sci. USA 77:7342-7346, 1980). Neutron activation analysis indicated that the nitrogenase 3 complex lacked significant amounts of Mo, V, Cr, Re, and W. Some Zn, however, was found in the dinitrogenase 3S and dinitrogenase 3F preparations. The pattern of substrate reduction efficiency was H+ greater than N2 greater than C2H2. The maximum specific activity found for N2 reduction was 38 nmol of NH3 per min per mg of protein (dinitrogenase 3S). Nitrogenase 3 was found to be extremely sensitive to O2, and activities could not be reproducibly maintained during freezing and thawing.  相似文献   

20.
The reduction of N2 to 2NH3 by Klebsiella pneumoniae nitrogenase was studied by a rapid-quench technique. The pre-steady-state time course for N2H4, formed on quenching by the acid-induced hydrolysis of an enzyme-bound intermediate in N2 reduction, showed a 230 ms lag followed by a damped oscillatory approach to a constant concentration in the steady state. The pre-steady-state time course for NH3 formation exhibited a lag of 500 ms and a burst phase that was essentially complete at 1.5s, before a steady-state rate was achieved. These time courses have been simulated by using a previously described kinetic model for the mechanism of nitrogenase action [Lowe & Thorneley (1984) Biochem. J. 224, 877-886]. A hydrazido(2-) structure (=N-NH2) is favoured for the intermediate that yields N2H4 on quenching. The NH3-formation data indicate enzyme-bound metallo-nitrido (identical to N) or -imido (=NH) intermediates formed after N-N bond cleavage to produce the first molecule of NH3 and which subsequently give the second molecule of NH3 by hydrolysis on quenching. The simulations require stoichiometric reduction of one N2 molecule at each Mo and the displacement of one H2 when N2 binds to the MoFe protein. Inhibition by H2 of N2-reduction activity occurs before the formation of the proposed hydrazido(2-) species, and is explained by H2 displacement of N2 at the active site.  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