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.
Nitrobacter agilis, which contains a very active nitrite dehydrogenase, was studied in vivo under anaerobic conditions by the 15N NMR technique. When incubated with equimolar 15NO3- and unlabeled nitrite (or 15NO2- and unlabeled nitrate) the bacterium catalyzed an isotope exchange reaction at rates about 10% those observed in the nitrite oxidase assay. When incubated with 18O-labeled 15NO2- and 18O-labeled 15NO3-, the 18O was observed to exchange at similar rates from both species into water. Finally, when incubated with equimolar [18O]nitrate and 15NO2-, intermolecular 18O transfer was observed to result in formation of double labeled nitrate and nitrite at similar rates. 18O was transferred from nitrate to a 15N species or to water at approximately equal rates under the conditions of the experiments. It is argued that the enzyme responsible for these exchange reactions is nitrite dehydrogenase and not nitrate reductase. This work and the related experiments of DiSpirito and Hooper (DiSpirito, A.A., and Hooper, A.B. (1986) J. Biol. Chem. 261, 10534-10537) represent the first demonstrations of intermolecular oxygen atom transfer among oxotransferases. Mechanistic implications are discussed.  相似文献   

2.
The effects of four species of denitrifying bacteria on the conversion of [15N]nitrite to trioxodinitrate (HN2O3-) and N2O and of trioxodinitrate to N2O were studied. For all species, the N2O produced in the presence of [15N]nitrite and trioxodinitrate was isotopically randomized throughout the period of incubation and was not composed at the outset predominantly of 14N2O or 14N2O plus 15N2O. The N2O produced was also heavily enriched in 15N at times when the trioxodinitrate pool was only weakly enriched in 15N. By 15N NMR, the N(2) position, but not the N(1) position, of trioxodinitrate was found to become progressively labeled with 15N during incubation with [15N]nitrite. These results argue that (a) the N-N bond of trioxodinitrate is not preserved in its conversion to N2O, (b) trioxodinitrate can be neither a free nor enzyme-bound intermediate in denitrifying bacteria, and (c) the pathways from nitrite and trioxodinitrate involve a common mononitrogen intermediate. The conclusion that this intermediate is probably nitroxyl (HNO), at least with Paracoccus denitrificans and Pseudomonas stutzeri, provides indirect evidence that N-N bond formation in denitrification can occur through the dimerization of nitroxyl.  相似文献   

3.
4.
The combined action of ammonia monooxygenase, AMO, (NH(3)+2e(-)+O(2)-->NH(2)OH) and hydroxylamine oxidoreductase, HAO, (NH(2)OH+H(2)O-->HNO(2)+4e(-)+4H(+)) accounts for ammonia oxidation in Nitrosomonas europaea. Pathways for electrons from HAO to O(2), nitrite, NO, H(2)O(2) or AMO are reviewed and some recent advances described. The membrane cytochrome c(M)552 is proposed to participate in the path between HAO and ubiquinone. A bc(1) complex is shown to mediate between ubiquinol and the terminal oxidase and is shown to be downstream of HAO. A novel, red, low-potential, periplasmic copper protein, nitrosocyanin, is introduced. Possible mechanisms for the inhibition of ammonia oxidation in cells by protonophores are summarized. Genes for nitrite- and NO-reductase but not N(2)O or nitrate reductase are present in the genome of Nitrosomonas. Nitrite reductase is not repressed by growth on O(2); the flux of nitrite reduction is controlled at the substrate level.  相似文献   

5.
Abstract Enumerations of nitrite-oxidizing bacteria in soil samples by a Most Probable Number technique, often showed relatively high cell numbers at a low nitrite concentration compared with the numbers of ammonium-oxidizing bacteria. It was hypothesized that the high numbers enumerated at low nitrite concentration would represent non-growing or organotrophically growing cells of nitrite-oxidizing species. In this paper, the sensitivity of non-growing Nitrobacter species to high nitrite concentrations as well as to low pH was examined. Different Nitrobacter species were pre-cultured at 0.5 mM nitrite. Non-growing cells differing in age were enumerated at different nitrite concentrations and pH values. The incubation period lasted for 5 months at 20°C. However, during the incubation periods of the older non-growing cells, it appeared that a period of 5 months might have been too short for reaching constant numbers. Early stationary cells of all species that were studied appeared not to be affected by high nitrite concentrations or low pH. Eight- and 18-month-old non-growing cells of Nitrobacter hamburgensis were also insensitive to 5 mM nitrite. The numbers of 8- and 18-month-old resting cells of N. vulgaris were only repressed by a combination of 5 mM nitrite and a low pH. Eight-month-old non-growing cells of N. winogradskyi were sensitive to 5 mM irrespective of pH, but 18-month-old cells only to 5 mM nitrate at low pH. The numbers of 8- and 18-month-old resting cells of N. winogradskyi serotype agilis were repressed by low pH rather than high nitrite concentration. Hence, it was concluded that the large differences in numbers of nitrite-oxidizing bacteria obtained with low and high nitrite concentrations in the incubation medium, was not likely to be due to the presence of non-growing Nitrobacter species in soil samples, but rather to the existence of organotrophically growing Nitrobacter cells.  相似文献   

6.
During the reaction of oxyhemoglobin (HbO2) with nitrite, the concentration of residual nitrite, nitrate, oxygen, and methemoglobin (Hb+) was determined successively. The results obtained at various pH values indicate the following stoichiometry for the overall reaction: 4HbO2 + 4NO2- 4H+ leads to 4Hb+ + 4NO3- + O2 + 2H2 O (Hb denotes hemoglobin monomer). NO2- binds with methemoglobin noncooperatively with a binding constant of 340 M-1 at pH 7.4 and 25 degrees C. Thus, the major part of Hb+ produced is aquomethemoglobin, not methemoglobin nitrite, when less than 2 equivalents of nitrite is used for the oxidation.  相似文献   

7.
The exchange of 18O between H218O and exogeneously added 15N16O?2 which occurs during oxidation of ammonia by Nitrosomonas is shown to occur one oxygen at a time. Conditions in which the exchange is diminished (notably the presence of 14NO2 and CCCP) allowed demonstration that water and dioxygen are each the source of one oxygen in nitrite produced from 15NH3. The nitrate produced in the presence of 18O2 consisted of 67 and 0% 15N18O16O? and 15N18O18O?, respectively. Analysis was made using the 18O-isotope shift in 15N-NMR.  相似文献   

8.
Soil emission of gaseous N oxides during nitrification of ammonium represents loss of an available plant nutrient and has an important impact on the chemistry of the atmosphere. We used selective inhibitors and a glucose amendment in a factorial design to determine the relative contributions of autotrophic ammonium oxidizers, autotrophic nitrite oxidizers, and heterotrophic nitrifiers to nitric oxide (NO) and nitrous oxide (N(2)O) emissions from aerobically incubated soil following the addition of 160 mg of N as ammonium sulfate kg. Without added C, peak NO emissions of 4 mug of N kg h were increased to 15 mug of N kg h by the addition of sodium chlorate, a nitrite oxidation inhibitor, but were reduced to 0.01 mug of N kg h in the presence of nitrapyrin [2-chloro-6-(trichloromethyl)-pyridine], an inhibitor of autotrophic ammonium oxidation. Carbon-amended soils had somewhat higher NO emission rates from these three treatments (6, 18, and 0.1 mug of N kg h after treatment with glucose, sodium chlorate, or nitrapyrin, respectively) until the glucose was exhausted but lower rates during the remainder of the incubation. Nitrous oxide emission levels exhibited trends similar to those observed for NO but were about 20 times lower. Periodic soil chemical analyses showed no increase in the nitrate concentration of soil treated with sodium chlorate until after the period of peak NO and N(2)O emissions; the nitrate concentration of soil treated with nitrapyrin remained unchanged throughout the incubation. These results suggest that chemoautotrophic ammonium-oxidizing bacteria are the predominant source of NO and N(2)O produced during nitrification in soil.  相似文献   

9.
Shen B  English AM 《Biochemistry》2005,44(42):14030-14044
Although biologically active, nitroxyl (HNO) remains one of the most poorly studied NO(x). Protein-based thiols are suspected targets of HNO, forming either a disulfide or sulfinamide (RSONH2) through an N-hydroxysulfenamide (RSNHOH) addition product. Electrospray ionization mass spectrometry (ESI-MS) is used here to examine the products formed during incubation of thiol proteins with the HNO donor, Angeli's salt (AS; Na2N2O3). Only the disulfide, cystine, was formed in incubates of 15 mM free Cys with equimolar AS at pH 7.0-7.4. In contrast, the thiol proteins (120-180 microM), human calbindin D(28k) (HCalB), glyceraldehyde-3-phosphate dehydrogenase (GAPDH), and bovine serum albumin (BSA) gave four distinct types of derivatives in incubates containing 0.9-2.5 mM AS. Ions at M + n x 31 units were detected in the ESI mass spectra of intact HCalB (n = 1-5) and GAPDH (n = 2), indicating conversion of thiol groups on these proteins to RSONH2 (+31 units). An ion at M + 14 dominated the mass spectrum of BSA, and intramolecular sulfinamide cross-linking of Cys34 to one of its neighboring Lys or Arg residues would account for this mass increase. Low abundant M + 14 adducts were observed for HCalB, which additionally formed mixed disulfides when free Cys was present in the AS incubates. Cys149 and Cys153 formed an intramolecular disulfide in the AS/GAPDH incubates. Since AS also produces nitrite above pH 5 (HN2O3(-) --> HNO + NO2(-)), incubation with NaNO2 served to confirm that protein modification was HNO-mediated, and prior blocking with the thiol-specific reagent, N-ethylmaleimide, demonstrated that thiols are the targets of HNO. The results provide the first systematic characterization of HNO-mediated derivatization of protein thiols.  相似文献   

10.
The mechanism of the denitrification and nitrosation reactions catalyzed by the heme cd-containing nitrite reductase from Pseudomonas stutzeri JM 300 has been studied with whole cell suspensions using H2(18)O, 15NO, and 15NO-2. The extent of H2(18)O exchange with the enzyme-bound nitrosyl intermediate, as determined by the 18O content of product N2O, decreased with increasing nitrite concentration, which is consistent with production of N2O by sequential reaction of two nitrite ions with the enzyme. Reaction of NO with whole cells in H2(18)O gave amounts of 18O in the N2O product consistent with equilibration of nitric oxide with a small pool of free nitrite. Using 15NO and NH2OH, competition between denitrification and nitrosation reactions was demonstrated, as is required if the enzyme-nitrosyl complex is an intermediate in both nitrosation and denitrification reactions. The first evidence for exchange of 18O between H2(18)O and a nitrosation intermediate occurring after the enzyme-nitrosyl complex, presumably an enzyme-bound nitrosamine, has been obtained. The collective results are most consistent with denitrification N2O originating via attack of NO-2 on a coordinated nitrosyl, as proposed earlier (Averill, B. A., and Tiedje, J. M. (1982) FEBS Lett. 138, 8-11).  相似文献   

11.
The nitrite oxidizing system of Nitrobacter winogradskyi   总被引:1,自引:0,他引:1  
Cytochrome components which participate in the oxidation of nitrite in Nitrobacter winogradskyi have been highly purified and their properties studied in detail. Cytochrome a1c1 is an iron-sulphur molybdoenzyme which has haems a and c and acts as a nitrite-cytochrome c oxidoreductase. Cytochrome c-550 is homologous to eukaryotic cytochrome c and acts as the electron mediator between cytochrome a1c1 and aa3-type cytochrome c oxidase. The oxidase is composed of two kinds of subunits, has two molecules of haem a and two atoms of copper in the molecule, and oxidizes actively eukaryotic ferrocytochrome c as well as its own ferrocytochrome c-550. Further, a flavoenzyme has been obtained which has transhydrogenase activity and catalyses reduction of NADP+ with benzylviologen radical. This enzyme may be responsible for production of NADPH in N. winogradskyi. The electron transfer against redox potential from NO2- to cytochrome c could be pushed through prompt removal by cytochrome aa3 of H+ formed by the dehydrogenation of NO2- + H2O. As cytochrome c in anaerobically kept cell-free extracts is rapidly reduced on addition of NO2-, a membrane potential does not seem necessary for the reduction of cytochrome c by cytochrome a1c1 with NO2- in vivo.  相似文献   

12.
Reduction of NO and NO2-by whole cells of eight strains of denitrifying bacteria known to contain either heme cd1 or copper-containing nitrite reductases (NiRs) has been examined in the presence of H218O. All organisms containing heme cd1 NiRs exhibited relatively large extents of exchange between NO2- and H218O (39-100%), as monitored by the 18O content of product N2O. Organisms containing copper NiRs gave highly variable results, with Achromobacter cycloclastes and Pseudomonas aureofaciens exhibiting no 18O incorporation and Rhodopseudomonas sphaeroides and Alcaligenes entrophus exhibiting complete exchange between NO2- and H218O. Organisms containing heme cd1 NiRs exhibited significant but lower levels of exchange between NO and H218O than between NO2- and H218O, while organisms containing copper NiRs gave significantly higher amounts of 18O incorporation than observed for the heme cd1 organisms. These results demonstrate the existence of an NO-derived species capable of undergoing O-atom exchange with H218O during the reduction of NO. Trapping experiments with 15NO, 14N3-, and crude extracts of R. sphaeroides support the electrophilic nature of this intermediate and suggest its formulation as an enzyme nitrosyl, E-NO+, analogous to that observed during reduction of NO2-. The observation of lower levels of 18O incorporation with NO2- than with NO as substrate for A. cycloclastes and P. aureofaciens indicates that, for these organisms at least, a sequential pathway involving free NO as an intermediate is significantly less important than a direct pathway in which N2O is formed via reaction of two NO2- ions on a single enzyme.  相似文献   

13.
Two related polytopic membrane proteins of the major facilitator family, NarK and NarU, catalyse nitrate uptake, nitrite export and nitrite uptake across the Escherichia coli cytoplasmic membrane by an unknown mechanism. A 12-helix model of NarU was constructed based upon six alkaline phosphatase and beta-galactosidase fusions to NarK and the predicted hydropathy for the NarK family. Fifteen residues conserved in the NarK-NarU protein family were substituted by site-directed mutagenesis, including four residues that are essential for nitrate uptake by Aspergillus nidulans: arginines Arg(87) and Arg(303) in helices 2 and 8, and two glycines in a nitrate signature motif. Despite the wide range of substitutions studied, in no case did mutation result in loss of one biochemical function without simultaneous loss of all other functions. A NarU+ NirC+ strain grew more rapidly and accumulated nitrite more rapidly than the isogenic NarU+ NirC(-) strain. Only the NirC+ strain consumed nitrite rapidly during the later stages of growth. Under conditions in which the rate of nitrite reduction was limited by the rate of nitrite uptake, NirC+ strains reduced nitrite up to 10 times more rapidly than isogenic NarU+ strains, indicating that both nitrite efflux and nitrite uptake are largely dependent on NirC. Isotope tracer experiments with [15N]nitrate and [14N]nitrite revealed that [15N]nitrite accumulated in the extracellular medium even when there was a net rate of nitrite uptake and reduction. We propose that NarU functions as a single channel for nitrate uptake and nitrite expulsion, either as a nitrate-nitrite antiporter, or more likely as a nitrate/H+ or nitrite/H+ channel.  相似文献   

14.
Nitrosomonas europaea and Nitrobacter winogradskyi (strain "Engel") were grown in ammonia-limited and nitrite-limited conditions, respectively, in a retentostat with complete biomass retention at 25 degrees C and pH 8. Fitting the retentostat biomass and oxygen consumption data of N. europaea and N. winogradskyi to the linear equation for substrate utilization resulted in up to eight-times-lower maintenance requirements compared to the maintenance energy demand (m) calculated from chemostat experiments. Independent of the growth rate at different stages of such a retention culture, the maximum specific oxygen consumption rate measured by mass spectrometric analysis of inlet and outlet gas oxygen content always amounted to approximately 45 micromol of O2 mg-1 of biomass-C x h-1 for both N. europaea and N. winogradskyi. When bacteria were starved for different time periods (up to 3 months), the spontaneous respiratory activity after an ammonia or nitrite pulse decreased with increasing duration of the previous starvation time period, but the observed decrease was many times faster for N. winogradskyi than for N. europaea. Likewise, the velocity of resuscitation decreased with extended time periods of starvation. The increase in oxygen consumption rates during resuscitation referred to the reviving population only, since in parallel no significant increase in the cell concentrations was detectable. N. europaea more readily recovers from starvation than N. winogradskyi, explaining the occasionally observed nitrite accumulation in the environment after ammonia becomes available. From chloramphenicol (100 microg x ml-1) inhibition experiments with N. winogradskyi, it has been concluded that energy-starved cells must have a lower protein turnover rate than nonstarved cells. As pointed out by Stein and Arp (L. Y. Stein and D. J. Arp, Appl. Environ. Microbiol. 64:1514-1521, 1998), nitrifying bacteria in soil have to cope with extremely low nutrient concentrations. Therefore, a chemostat is probably not a suitable tool for studying their physiological properties during a long-lasting nutrient shortage. In comparison with chemostats, retentostats offer a more realistic approach with respect to substrate provision and availability.  相似文献   

15.
The alphaproteobacterium Nitrobacter winogradskyi (ATCC 25391) is a gram-negative facultative chemolithoautotroph capable of extracting energy from the oxidation of nitrite to nitrate. Sequencing and analysis of its genome revealed a single circular chromosome of 3,402,093 bp encoding 3,143 predicted proteins. There were extensive similarities to genes in two alphaproteobacteria, Bradyrhizobium japonicum USDA110 (1,300 genes) and Rhodopseudomonas palustris CGA009 CG (815 genes). Genes encoding pathways for known modes of chemolithotrophic and chemoorganotrophic growth were identified. Genes encoding multiple enzymes involved in anapleurotic reactions centered on C2 to C4 metabolism, including a glyoxylate bypass, were annotated. The inability of N. winogradskyi to grow on C6 molecules is consistent with the genome sequence, which lacks genes for complete Embden-Meyerhof and Entner-Doudoroff pathways, and active uptake of sugars. Two gene copies of the nitrite oxidoreductase, type I ribulose-1,5-bisphosphate carboxylase/oxygenase, cytochrome c oxidase, and gene homologs encoding an aerobic-type carbon monoxide dehydrogenase were present. Similarity of nitrite oxidoreductases to respiratory nitrate reductases was confirmed. Approximately 10% of the N. winogradskyi genome codes for genes involved in transport and secretion, including the presence of transporters for various organic-nitrogen molecules. The N. winogradskyi genome provides new insight into the phylogenetic identity and physiological capabilities of nitrite-oxidizing bacteria. The genome will serve as a model to study the cellular and molecular processes that control nitrite oxidation and its interaction with other nitrogen-cycling processes.  相似文献   

16.
Some recent studies on the pathway of nitrogen and the reductases of denitrification are reviewed. The available evidence suggests that while the intermediates of denitrification can remain enzyme-bound (presumably to nitrite reductase) prior to formation of N2O, NO and nitroxyl (HNO) can be released in part by certain bacteria. Release of NO is recognized by a nitrite/NO?15N exchange reaction and isotopic scrambling in product N2O; release of nitroxyl by Pseudomonas stutzeri is recognized by isotopic scrambling of nitrite and NO in product N2O in absence of exchange and affords evidence that the first N?N bond forms in denitrification at the N1+ redox level. The recent purification and partial characterization of nitrous oxide reductase are described. The ability of the dissimilatory nitrite reductase to activate nitrite for nitrosyl transfer affords a new chemical probe into the mechanism of action of this central enzyme. It would appear that reduction of nitrite is subject to electrophilic catalysis. 18O studies show that dissociation of nitrite from nitrite reductase can be slow relative to competing reduction or nitrosyl transfer.  相似文献   

17.
Kimura Y  Ishii A  Yamanari T  Ono TA 《Biochemistry》2005,44(21):7613-7622
In photosynthetic water oxidation, two water molecules are converted to an oxygen molecule through five reaction intermediates, designated S(n) (n = 0-4), at the catalytic Mn cluster of photosystem II. To understand the mechanism of water oxidation, changes in the chemical nature of the substrate water as well as the Mn cluster need to be defined during S-state cycling. Here, we report for the first time a complete set of Fourier transform infrared difference spectra during S-state cycling in the low-frequency (670-350 cm(-1)) region, in which interactions between the Mn cluster and its ligands can be detected directly, in PS II core particles from Thermosynechococcus elongatus. Furthermore, vibrations from oxygen and/or hydrogen derived from the substrate water and changes in them during S-state cycling were identified using multiplex isotope-labeled water, including H2(18)O, D2(16)O, and D2(18)O. Each water isotope affected the low-frequency S-state cycling spectra, characteristically. The bands sensitive only to (16)O/(18)O exchange were assigned to the modes from structures involving Mn and oxygen having no interactions with hydrogen, while the bands sensitive only to H/D exchange were assigned to modes from amino acid side chains and/or polypeptide backbones that associate with water hydrogen. The bands sensitive to both (16)O/(18)O and H/D exchanges were attributed to the structure involving Mn and oxygen structurally coupled with hydrogen in a direct or an indirect manner through hydrogen bonds. These bands include the changes of intermediate species derived from substrate water during the process of photosynthetic water oxidation.  相似文献   

18.
We screened actinomycete strains for dinitrogen (N(2))-producing activity and discovered that Streptomyces antibioticus B-546 evolves N(2) and some nitrous oxide (N(2)O) from nitrate (NO(3)(-)). Most of the N(2) that evolved from the heavy isotope ([(15)N]NO(3)(-)) was (15)N(14)N, indicating that this nitrogen species consists of two atoms, one arising from NO(3)(-) and the other from different sources. This phenomenon is similar to codenitrification in fungi. The strain also evolved less, but significant, amounts of (15)N(15)N from [(15)N]NO(3)(-) in addition to (15)N(15)NO with concomitant cell growth. Prior to the production of N(2) and N(2)O, NO(3)(-) was rapidly reduced to nitrite (NO(2)(-)) accompanied by distinct cell growth, showing that the actinomycete strain is a facultative anaerobe that depends on denitrification and nitrate respiration for anoxic growth. The cell-free activities of denitrifying enzymes could be reconstituted, supporting the notion that the (15)N(15)N and (15)N(15)NO species are produced by denitrification from NO(3)(-) via NO(2)(-). We therefore demonstrated a unique system in an actinomycete that produces gaseous nitrogen (N(2) and N(2)O) through both denitrification and codenitrification. The predominance of codenitrification over denitrification along with oxygen tolerance is the key feature of nitrate metabolism in this actinomycete.  相似文献   

19.
The intramolecular distribution of nitrogen isotopes in N2O is an emerging tool for defining the relative importance of microbial sources of this greenhouse gas. The application of intramolecular isotopic distributions to evaluate the origins of N2O, however, requires a foundation in laboratory experiments in which individual production pathways can be isolated. Here we evaluate the site preferences of N2O produced during hydroxylamine oxidation by ammonia oxidizers and by a methanotroph, ammonia oxidation by a nitrifier, nitrite reduction during nitrifier denitrification, and nitrate and nitrite reduction by denitrifiers. The site preferences produced during hydroxylamine oxidation were 33.5 +/- 1.2 per thousand, 32.5 +/- 0.6 per thousand, and 35.6 +/- 1.4 per thousand for Nitrosomonas europaea, Nitrosospira multiformis, and Methylosinus trichosporium, respectively, indicating similar site preferences for methane and ammonia oxidizers. The site preference of N2O from ammonia oxidation by N. europaea (31.4 +/- 4.2 per thousand) was similar to that produced during hydroxylamine oxidation (33.5 +/- 1.2 per thousand) and distinct from that produced during nitrifier denitrification by N. multiformis (0.1 +/- 1.7 per thousand), indicating that isotopomers differentiate between nitrification and nitrifier denitrification. The site preferences of N2O produced during nitrite reduction by the denitrifiers Pseudomonas chlororaphis and Pseudomonas aureofaciens (-0.6 +/- 1.9 per thousand and -0.5 +/- 1.9 per thousand, respectively) were similar to those during nitrate reduction (-0.5 +/- 1.9 per thousand and -0.5 +/- 0.6 per thousand, respectively), indicating no influence of either substrate on site preference. Site preferences of approximately 33 per thousand and approximately 0 per thousand are characteristic of nitrification and denitrification, respectively, and provide a basis to quantitatively apportion N2O.  相似文献   

20.
Five Tn5 mutants of Pseudomonas fluorescens AK-15 deficient in dissimilatory reduction of nitrite were isolated and characterized. Two insertions occurred inside the nitrite reductase structural gene (nirS) and resulted in no detectable nitrite reductase protein on a Western immunoblot. One mutant had Tn5 inserted inside nirC, the third gene in the same operon, and produced a defective nitrite reductase protein. Two other mutants had insertions outside of this nir operon and also produced defective proteins. All of the Nir- mutants characterized showed not only loss of nitrite reductase activity but also a significant decrease in nitric oxide reductase activity. When cells were incubated with 15NO in H2(18)O, about 25% of the oxygen found in nitrous oxide exchanged with H2O. The extent of exchange remained constant throughout the reaction, indicating the incorporation of 18O from H2(18)O reached equilibrium rapidly. In all nitrite reduction-deficient mutants, less than 4% of the 18O exchange was found, suggesting that the hydration and dehydration step was altered. These results indicate that the factors involved in dissimilatory reduction of nitrite influenced the subsequent NO reduction in this organism.  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