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 281 毫秒
1.
The Azospirillum brasilense Sp245 napABC genes, encoding nitrate reductase activity, were isolated and sequenced. The derived protein sequences are very similar throughout the whole Nap segment to the NapABC protein sequences of Escherichia coli, Pseudomonas sp. G-179, Ralstonia eutropha, Rhodobacter sphaeroides, and Paracoccus denitrificans. Based on whole-cell nitrate reductase assays with the artificial electron donors benzyl viologen and methyl viologen, and assays with periplasmic cell-free extracts, it was concluded that the napABC-encoded enzyme activity in Azospirillum brasilense Sp245 corresponds to a periplasmic dissimilatory nitrate reductase, which was expressed under anoxic conditions and oxic conditions. A kanamycin-resistant Azospirillum brasilense Sp245 napA insertion mutant was constructed. The mutant still expressed assimilatory nitrate reductase activity, but was devoid of its periplasmic dissimilatory nitrate reductase activity.  相似文献   

2.
Preliminary studies showed that the periplasmic nitrate reductase (Nap) of Rhodobacter sphaeroides and the membrane-bound nitrate reductases of Escherichia coli are able to reduce selenate and tellurite in vitro with benzyl viologen as an electron donor. In the present study, we found that this is a general feature of denitrifiers. Both the periplasmic and membrane-bound nitrate reductases of Ralstonia eutropha, Paracoccus denitrificans, and Paracoccus pantotrophus can utilize potassium selenate and potassium tellurite as electron acceptors. In order to characterize these reactions, the periplasmic nitrate reductase of R. sphaeroides f. sp. denitrificans IL106 was histidine tagged and purified. The V(max) and K(m) were determined for nitrate, tellurite, and selenate. For nitrate, values of 39 micromol x min(-1) x mg(-1) and 0.12 mM were obtained for V(max) and K(m), respectively, whereas the V(max) values for tellurite and selenate were 40- and 140-fold lower, respectively. These low activities can explain the observation that depletion of the nitrate reductase in R. sphaeroides does not modify the MIC of tellurite for this organism.  相似文献   

3.
The periplasmic dissimilatory nitrate reductase from Rhodobacter capsulatus N22DNAR+ has been purified. It comprises a single type of polypeptide chain with subunit molecular weight 90,000 and does not contain heme. Chlorate is not an alternative substrate. A molybdenum cofactor, of the pterin type found in both nitrate reductases and molybdoenzymes from various sources, is present in nitrate reductase from R. capsulatus at an approximate stoichiometry of 1 molecule per polypeptide chain. This is the first report of the occurrence of the cofactor in a periplasmic enzyme. Trimethylamine-N-oxide reductase activity was fractionated by ion exchange chromatography of periplasmic proteins. The fractionated material was active towards dimethylsulphoxide, chlorate and methionine sulphoxide, but not nitrate. A catalytic polypeptide of molecular weight 46,000 was identified by staining for trimethylamine-N-oxide reductase activity after polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulphate. The same polypeptide also stained for dimethylsulphoxide reductase activity which indicates that trimethylamine-N-oxide and dimethylsulphoxide share a common reductase.Abbreviations DMSO dimethylsulphoxide - LDS lithium dodecyl sulphate - MVH reduced methylviologen - PAGE polyacrylamide gel electrophoresis - SDS sodium dodecyl sulphate - TMAO trimethylamine-N-oxide  相似文献   

4.
The periplasmic nitrate reductase (NapAB), a member of the DMSO reductase superfamily, catalyzes the first step of the denitrification process in bacteria. In this heterodimer, a di-heme NapB subunit is associated to the catalytic NapA subunit that binds a [4Fe-4S] cluster and a bis(molybdopterin guanine dinucleotide) cofactor. Here, we report the kinetic characterization of purified mutated heterodimers from Rhodobacter sphaeroides. By combining site-directed mutagenesis, redox potentiometry, EPR spectroscopy, and enzymatic characterization, we investigate the catalytic role of two conserved residues (M153 and R392) located in the vicinity of the molybdenum active site. We demonstrate that M153 and R392 are involved in nitrate binding: the Vm measured on the M153A and R392A mutants are similar to that measured on the wild-type enzyme, whereas the Km for nitrate is increased 10-fold and 200-fold, respectively. The use of an alternative enzymatic assay led us to discover that NapAB is uncompetitively inhibited by Zn2+ ions (Ki' = 1 microM). We used this property to further probe the active site access in the mutant enzymes. It is proposed that R392 acts as a filter by preventing a direct reduction of the Mo atom by small reducing molecules and partially protecting the active site against zinc inhibition. In addition, we show that M153 is a key residue mediating this inhibition likely by coordinating Zn2+ ions via its sulfur atom. This residue is not conserved in the DMSO reductase superfamily while it is conserved in the periplasmic nitrate reductase family. Zinc inhibition is therefore likely to be specific and restricted to periplasmic nitrate reductases.  相似文献   

5.
Alcaligenes eutrophus H16 shows three distinct nitrate reductase activities (U. Warnecke-Eberz and B. Friedrich, Arch. Microbiol. 159:405-409, 1993). The periplasmic enzyme, designated NAP (nitrate reductase, periplasmic), has been isolated. The 80-fold-purified heterodimeric enzyme catalyzed nitrate reduction with reduced viologen dyes as electron donors. The nap genes were identified in a library of A. eutrophus H16 megaplasmid DNA by using oligonucleotide probes based on the amino-terminal polypeptide sequences of the two NAP subunits. The two structural genes, designated napA and napB, code for polypeptides of 93 and 18.9 kDa, respectively. Sequence comparisons indicate that the putative gene products are translated with signal peptides of 28 and 35 amino acids, respectively. This is compatible with the fact that NAP activity was found in the soluble fraction of cell extracts and suggests that the mature enzyme is located in the periplasm. The deduced sequence of the large subunit, NAPA, contained two conserved amino-terminal stretches of amino acids found in molybdenum-dependent proteins such as nitrate reductases and formate dehydrogenases, suggesting that NAPA contains the catalytic site. The predicted sequence of the small subunit, NAPB, revealed two potential heme c-binding sites, indicating its involvement in the transfer of electrons. An insertion in the napA gene led to a complete loss of NAP activity but did not abolish the ability of A. eutrophus to use nitrate as a nitrogen source or as an electron acceptor in anaerobic respiration. Nevertheless, the NAP-deficient mutant showed delayed growth after transition from aerobic to anaerobic respiration, suggesting a role for NAP in the adaptation to anaerobic metabolism.  相似文献   

6.
The Pseudomonas fluorescens YT101 gene narG, which encodes the catalytic alpha subunit of the respiratory nitrate reductase, was disrupted by insertion of a gentamicin resistance cassette. In the Nar(-) mutants, nitrate reductase activity was not detectable under all the conditions tested, suggesting that P. fluorescens YT101 contains only one membrane-bound nitrate reductase and no periplasmic nitrate reductase. Whereas N(2)O respiration was not affected, anaerobic growth with NO(2) as the sole electron acceptor was delayed for all of the Nar(-) mutants following a transfer from oxic to anoxic conditions. These results provide the first demonstration of a regulatory link between nitrate and nitrite respiration in the denitrifying pathway.  相似文献   

7.
A strain of Pseudomonas putida that can express a nitrate reductase that is located in the periplasmic compartment was isolated from freshwater. The enzyme was active in vivo during arginine fermentation and at the onset of oxygen limitation in batch cultures. The activity of the enzyme increased the yield of bacteria following fermentative growth under anoxic conditions with arginine, but nitrate reduction did not support growth on nonfermentable carbon substrates under anoxic conditions. Cells expressing the periplasmic nitrate reductase were capable of reducing nitrate in the presence of oxygen. Nitrate reduction under oxic conditions was clearly coupled to a respiratory electron transport chain because: (1) the process was sensitive to the respiratory inhibitors rotenone and 2-n-heptyl-4-hydroxyquinoline N-oxide, and (2) membrane-bound and periplasmic cytochromes were involved. This is the first report of the presence of a periplasmic nitrate reductase in a member of the proteobacteria.  相似文献   

8.
Seven genes, napKEFDABC, encoding the periplasmic nitrate reductase system were cloned from the denitrifying phototrophic bacterium Rhodobacter sphaeroides f. sp. denitrificans IL106. Two transmembrane proteins, NapK and NapE, an iron-sulfur protein NapF, a soluble protein NapD, a catalytic subunit of nitrate reductase precursor NapA, a soluble c-type diheme cytochrome precursor NapB, and a membrane-anchored c-type tetraheme cytochrome NapC were deduced as the gene products. Every mutant in which each nap gene was disrupted by omega-cassette insertion lost nitrate reductase activity as well as the ability of cells to grow with nitrate under anaerobic-dark conditions. A transconjugant of the napD-disrupted mutant with a plasmid bearing the napKEFDABC genes recovered both nitrate reductase activity and nitrate-dependent anaerobic-dark growth of cells. Denitrification activity, which was not observed in the napD mutant, was also restored by the conjugation. These results indicate that the periplasmic nitrate reductase encoded by the napKEFDABC genes is the enzyme responsible for denitrification in this phototroph, although the presence of a membrane-bound nitrate reductase has been reported in the same strain.  相似文献   

9.
Molybdenum enzymes containing the pterin cofactor are a diverse group of enzymes that catalyse in general oxygen atom transfer reactions. Aiming at studying the amino acid residues, which are important for the enzymatic specificity, we used nitrate reductase from Ralstonia eutropha (R.e.NAP) as a model system for mutational studies at the active site. We mutated amino acids at the Mo active site (Cys181 and Arg421) as well as amino acids in the funnel leading to it (Met182, Asp196, Glu197, and the double mutant Glu197-Asp196). The mutations were made on the basis of the structural comparison of nitrate reductases with formate dehydrogenases (FDH), which show very similar three-dimensional structures, but clear differences in amino acids surrounding the active site. For mutations Arg421Lys and Glu197Ala we found a reduced nitrate activity while the other mutations resulted in complete loss of activity. In spite of the partial of total loss of nitrate reductase activity, these mutants do not, however, display FDH activity.  相似文献   

10.
A Paracoccus denitrificans strain (M6Ω) unable to use nitrate as a terminal electron acceptor was constructed by insertional inactivation of the periplasmic and membrane-bound nitrate reductases. The mutant strain was able to grow aerobically with nitrate as the sole nitrogen source. It also grew anaerobically with nitrate as sole nitrogen source when nitrous oxide was provided as a respiratory electron acceptor. These growth characteristics are attributed to the presence of a third, assimilatory nitrate reductase. Nitrate reductase activity was detectable in intact cells and soluble fractions using nonphysiological electron donors. The enzyme activity was not detectable when ammonium was included in the growth medium. The results provide an unequivocal demonstration that P. denitrificans can express an assimilatory nitrate reductase in addition to the well-characterised periplasmic and membrane-bound nitrate reductases. Received: 12 August 1996 / Accepted: 29 October 1996  相似文献   

11.
Cyanobacterial NAD(P)(+)-reducing reversible hydrogenases comprise five subunits. Four of them (HoxF, HoxU, HoxY, and HoxH) are also found in the well-described related enzyme from Ralstonia eutropha. The fifth one (HoxE) is not encoded in the R. eutropha genome, but shares homology with the N-terminal part of R. eutropha HoxF. However, in cyanobacteria, HoxE contains a 2Fe-2S cluster-binding motif that is not found in the related R. eutropha sequence. In order to obtain some insights into the role of HoxE in cyanobacteria, we deleted this subunit in Synechocystis PCC6803. Three types of interaction of the cyanobacterial hydrogenase with pyridine nucleotides were tested: (a) reductive activation of the NiFe site, for which NADPH was found to be more efficient than NADH; (b) H(2) production, for which NADH appeared to be a more efficient electron donor than NADPH; and (c) H(2) oxidation, for which NAD(+) was a much better electron acceptor than NADP(+). Upon hoxE deletion, the Synechocystis hydrogenase active site remained functional with artificial electron donors or acceptors, but the enzyme became unable to catalyze H(2) production or uptake with NADH/NAD(+). However, activation of the electron transfer-independent H/D exchange reaction by NADPH was still observed in the absence of HoxE, whereas activation of this reaction by NADH was lost. These data suggest different mechanisms for diaphorase-mediated electron donation and catalytic site activation in cyanobacterial hydrogenase.  相似文献   

12.
Archaeoglobus fulgidus, a hyperthermophilic sulfate-reducing Archaeon, contains high Fe(3+)-EDTA reductase activity in its soluble protein fraction. The corresponding enzyme, which constitutes about 0.75% of the soluble protein, was purified 175-fold to homogeneity. Based on SDS-polyacrylamide gel electrophoresis, the ferric reductase consists of a single subunit with a M(r) of 18,000. The M(r) of the native enzyme was determined by size exclusion chromatography to be 40,000 suggesting that the native ferric reductase is a homodimer. The enzyme uses both NADH and NADPH as electron donors to reduce Fe(3+)-EDTA. Other Fe(3+) complexes and dichlorophenolindophenol serve as alternative electron acceptors, but uncomplexed Fe(3+) is not utilized. The purified enzyme strictly requires FMN or FAD as a catalytic intermediate for Fe(3+) reduction. Ferric reductase also reduces FMN and FAD, but not riboflavin, with NAD(P)H which classifies the enzyme as a NAD(P)H:flavin oxidoreductase. The enzyme exhibits a temperature optimum of 88 degrees C. When incubated at 85 degrees C, the enzyme activity half-life was 2 h. N-terminal sequence analysis of the purified ferric reductase resulted in the identification of the hypothetical gene, AF0830, of the A. fulgidus genomic sequence. The A. fulgidus ferric reductase shares amino acid sequence similarity with a family of NAD(P)H:FMN oxidoreductases but not with any ferric reductases suggesting that the A. fulgidus ferric reductase is a novel enzyme.  相似文献   

13.
The frequent abundance of arsenic in the environment has guided the evolution of enzymes for the reduction of arsenate. The arsenate reductases (ArsC) from different sources have unrelated sequences and structural folds, and can be divided into different classes on the basis of their structures, reduction mechanisms and the locations of catalytic cysteine residues. The thioredoxin-coupled arsenate reductase class is represented by Staphylococcus aureus pI258 ArsC and Bacillus subtilis ArsC. The ArsC from Escherichia coli plasmid R773 and the eukaryotic ACR2p reductase from Saccharomyces cerevisiae represent two distinct glutaredoxin-linked ArsC classes. All are small cytoplasmic redox enzymes that reduce arsenate to arsenite by the sequential involvement of three different thiolate nucleophiles that function as a redox cascade. In contrast, the ArrAB complex is a bacterial heterodimeric periplasmic or a surface-anchored arsenate reductase that functions as a terminal electron acceptor and transfers electrons from the membrane respiratory chain to arsenate. Finally, the less well documented arsenate reductase activity of the monomeric arsenic(III) methylase, which is an S-adenosylmethionine (AdoMet)-dependent methyltransferase. After each oxidative methylation cycle and before the next methylation step, As(V) is reduced to As(III). Methylation by this enzyme is also considered an arsenic-resistance mechanism for bacteria, fungi and mammals.  相似文献   

14.
Dissimilatory nitrite reductase catalyses the reduction of nitrite (NO(2)(-)) to nitric oxide (NO). Copper-containing nitrite reductases contain both type 1 and type 2 Cu sites. Electron transfer from redox partners is presumed to be mediated via the type 1 Cu site and used at the catalytic type 2 Cu centre along with the substrate nitrite. At the type 2 Cu site, Asp92 has been identified as a key residue in substrate utilisation, since it hydrogen bonds to the water molecule at the nitrite binding site. We have also suggested that protons enter the catalytic site via Asp92, through a water network that is mediated by His254. The role of these residues has been investigated in the blue copper nitrite reductase from Alcaligenes xylosoxidans (NCIMB 11015) by a combination of point mutation, enzymatic activity measurement and structure determination.In addition, it has been suggested that the enzyme operates via an ordered mechanism where an electron is transferred to the type 2 Cu site largely when the second substrate nitrite is bound and that this is controlled via the lowering of the redox potential of the type 2 site when it is loaded with nitrite. Thus, a small perturbation of the type 1 Cu site should result in a significant effect on the activity of the enzyme. For this reason a mutation of Met144, which is the weakest ligand of the type 1 Cu, is investigated. The structures of H254F, D92N and M144A have been determined to 1.85 A, 1.9 A and 2.2 A resolution, respectively. The D92N and H254F mutants have negligible or no activity, while the M144A mutant has 30 % activity of the native enzyme. Structural and spectroscopic data show that the loss of activity in H254F is due to the catalytic site being occupied by Zn while the loss/reduction of activity in D92N/M144A are due to structural reasons. The D92N mutation results in the loss of the Asp92 hydrogen bond to the Cu-ligated water. Therefore, the ligand is no longer able to perform proton abstraction. Even though the loss of activity in H254F is due to lack of catalytic Cu, the mutation does cause the disruption of the water network, confirming its key role in proton channel. The structure of the H254F mutant is the first case where full occupancy Zn at the type 2 Cu site is observed, but despite the previously noted similarity of this site to the carbonic anhydrase catalytic site, no carbonic anhydrase activity is observed. The H254F and D92N mutant structures provide, for the first time, observation of surface Zn sites which may act as a Zn sink and prevent binding of Zn at the catalytic Cu site in the native enzyme.  相似文献   

15.
1. The respiratory nitrate reductase of Klebsiella aerogenes was solubilized from the bacterial membranes by deoxycholate and purified further by means of gel chromatography in the presence of deoxycholate, and anion-exchange chromatography. 2. Dependent on the isolation procedure two different homogeneous forms of the enzyme, having different subunit compositions, can be obtained. These forms are designated nitrate reductase I and nitrate reductase II. Both enzyme preparations are isolated as tetramers having sedimentation constants (s20,w) of 22.1 S and 21.7 S for nitrate reductase I and II, respectively. The nitrate reductase I tetramer has a molecular weight of about 106. 3. In the presence of deoxycholate both enzyme preparations dissociate reversibly into their respective monomeric forms. The monomeric form of nitrate reductase I has a molecular weight of about 260 000 and a sedimentation constant of 9.8 S. For nitrate reductase II these values are 180 000 and 8.5 S, respectively. 4. Nitrate reductase I consists of three different subunits, having molecular weights of 117 000; 57 000 and 52 000, which are present in a 1:1:2 molar ratio, respectively. Nitrate reductase II contains only the subunits with a molecular weight of 117 000 and 57 000 in a equimolar ratio. 5. Treatment at pH 9.5 in the presence of deoxycholate and 0.05 M NaCl or ageing removes the 52 000 Mr subunit from nitrate reductase I. This smallest subunit, in contrast to the other subunits, is a basic protein. 6. The 52 000 Mr subunit has no catalytic function in the intramolecular electron transfer from reduced benzylviologen to nitrate. However, it appears to have a structural function since nitrate reductase II, which lacks this subunit, is much more labile than nitrate reductase I. Inactivation of nitrate reductase II can be prevented by the presence of deoxycholate. 7. The spectrum of the enzyme resembles that of iron-sulfur proteins. No cytochromes or contaminating enzyme activities are present in the purified enzyme. Only reduced benzylviologen was found to be capable of acting as an electron donor. 8. p-Chlormercuribenzoate enhances the enzymatic activity at concentrations of 0.1 mM and lower. At higher p-chlormercuribenzoate concentrations the enzymatic activity is inhibited non-competitively with either nitrate or benzylviologen as a substrate. The inhibition is not counteracted by cysteine.  相似文献   

16.
The structure of the respiratory nitrate reductase (NapAB) from Rhodobacter sphaeroides, the periplasmic heterodimeric enzyme responsible for the first step in the denitrification process, has been determined at a resolution of 3.2 A. The di-heme electron transfer small subunit NapB binds to the large subunit with heme II in close proximity to the [4Fe-4S] cluster of NapA. A total of 57 residues at the N- and C-terminal extremities of NapB adopt an extended conformation, embracing the NapA subunit and largely contributing to the total area of 5,900 A(2) buried in the complex. Complex formation was studied further by measuring the variation of the redox potentials of all the cofactors upon binding. The marked effects observed are interpreted in light of the three-dimensional structure and depict a plasticity that contributes to an efficient electron transfer in the complex from the heme I of NapB to the molybdenum catalytic site of NapA.  相似文献   

17.
Three subfamilies of metallopeptidase family M16 enzymes—M16A, M16B, and M16C—are widely distributed among eukaryotes and prokaryotes. SPH2681, a periplasmic M16B protein found in Sphingomonas sp. strain A1, contains an HXXEH motif essential for Zn2+ binding and catalytic activity. SPH2682 is another member of M16B, which lacks the metal-binding motif but conserves an active-site R/Y pair commonly found in the C-terminal half of M16 enzymes. Two genes coding for SPH2681 and SPH2682 assemble into a single operon in the bacterial genome. This study determined SPH2681 to be constitutively expressed in strain A1 cells grown on different carbon sources, suggesting a more general cellular function. SPH2681 and SPH2681/SPH2682 were overexpressed in Escherichia coli, purified, and characterized. SPH2681 was found to associate with SPH2682, forming a heterosubunit enzyme with peptidase activity, while SPH2681 alone exhibited no enzymatic activity. X-ray crystallography of the SPH2681/SPH2682 complex revealed two conformations (open and closed heterodimeric forms) within the same crystal. Compared with the closed form, the open form contains two subunits rotated away from each other by approximately 8°, increasing the distance between the zinc ion and active-site residues by up to 8 Å. In addition, many hydrogen bonds are formed or broken on change between the conformations of the heterodimers, suggesting that subunit dynamics is a prerequisite for catalysis. To our knowledge, this is the first report on both conformational forms of the same M16 peptidase, providing a unique insight into the general proteolytic mechanism of M16 proteases.  相似文献   

18.
Acetohydroxyacid synthase (AHAS; EC 2.2.1.6) is a thiamin diphosphate- (ThDP)- and FAD-dependent enzyme that catalyzes the first common step in the biosynthetic pathway of the branched-amino acids (BCAAs) leucine, isoleucine, and valine. The gene from Haemophilus influenzae that encodes the AHAS catalytic subunit was cloned, overexpressed in Escherichia coli BL21(DE3), and purified to homogeneity. The purified H. influenzae AHAS catalytic subunit (Hin-AHAS) appeared as a single band on SDS-PAGE gel, with a molecular mass of approximately 63 kDa. The enzyme catalyzes the condensation of two molecules of pyruvate to form acetolactate, with a K(m) of 9.2mM and the specific activity of 1.5 micromol/min/mg. The cofactor activation constant (K(c)=13.5 microM) and the dissociation constant (K(d)=3.3 microM) of ThDP were also determined by enzymatic assay and tryptophan fluorescence quenching studies, respectively. We screened a chemical library to discover new inhibitors of the Hin AHAS catalytic subunit. Through which, AVS-2087 (IC(50)=0.53 microM), KSW30191 (IC(50)=1.42 microM), and KHG20612 (IC(50)=4.91 microM) displayed potent inhibition as compare to sulfometuron methyl (IC(50)=276.31 microM).  相似文献   

19.
Summary The nucleotide sequence of the gene cysH from Escherichia coli K12 was determined. The open reading frame was 735 nucleotides in length; it was flanked by a repetitive palindromic sequence centred 36 nucleotides upstream of cysH and a terminator-like structure located 20 nucleotides downstream. CysH encoded a colypeptide of Mr 27927 consisting of 244 amino acids. The gene product was isolated as a homodimer exhibiting phospo-adenylylsulphate reductase (PAPS reductase) activity. The active enzyme was devoid of electron transferring cofactors and contained only one cysteine per subunit. Reduction of the enzyme by dithiols resulted in a shift of the apparent molecular weight from 44000 to 62000 without formation of an enzyme-thioredoxin complex.
  相似文献   

20.
Periplasmic nitrate reductase (NapABC enzyme) has been characterized from a variety of proteobacteria, especially Paracoccus pantotrophus. Whole-genome sequencing of Escherichia coli revealed the structural genes napFDAGHBC, which encode NapABC enzyme and associated electron transfer components. E. coli also expresses two membrane-bound proton-translocating nitrate reductases, encoded by the narGHJI and narZYWV operons. We measured reduced viologen-dependent nitrate reductase activity in a series of strains with combinations of nar and nap null alleles. The napF operon-encoded nitrate reductase activity was not sensitive to azide, as shown previously for the P. pantotrophus NapA enzyme. A strain carrying null alleles of narG and narZ grew exponentially on glycerol with nitrate as the respiratory oxidant (anaerobic respiration), whereas a strain also carrying a null allele of napA did not. By contrast, the presence of napA+ had no influence on the more rapid growth of narG+ strains. These results indicate that periplasmic nitrate reductase, like fumarate reductase, can function in anaerobic respiration but does not constitute a site for generating proton motive force. The time course of phi(napF-lacZ) expression during growth in batch culture displayed a complex pattern in response to the dynamic nitrate/nitrite ratio. Our results are consistent with the observation that phi(napF-lacZ) is expressed preferentially at relatively low nitrate concentrations in continuous cultures (H. Wang, C.-P. Tseng, and R. P. Gunsalus, J. Bacteriol. 181:5303-5308, 1999). This finding and other considerations support the hypothesis that NapABC enzyme may function in E. coli when low nitrate concentrations limit the bioenergetic efficiency of nitrate respiration via NarGHI 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