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.
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   

2.
Zostera marina L. is an angiosperm that grows in a medium in which inorganic phosphate (P(i)) and nitrate (NO(3)(-)) are present in micromolar concentrations and must be absorbed against a steep electrochemical potential gradient. The operation of a Na(+)-dependent NO(3)(-) transport was previously demonstrated in leaf cells of this plant, suggesting that other Na(+)-coupled systems could mediate the uptake of anions. To address this question, P(i) transport was studied in leaves and roots of Z. marina, as well as NO(3)(-) uptake in roots. Electrophysiological studies demonstrated that micromolar concentrations of P(i) induced depolarizations of the plasma membrane of root cells. However, this effect was not observed in leaf cells. P(i)-induced depolarizations showed Michaelis-Menten kinetics (K(m)=1.5+/-0.6 microM P(i); D(max)=7.8+/-0.8 mV), and were not observed in the absence of Na(+). However, depolarizations were restored when Na(+) was resupplied. NO(3)(-) additions also evoked depolarizations of the plasma membrane of root cells only in the presence of Na(+). Both NO(3)(-)- and P(i)-induced depolarizations were accompanied by an increase in cytoplasmic Na(+) activity, detected by Na(+)-sensitive microelectrodes. P(i) net uptake (measured in depletion experiments) was stimulated by Na(+). These results strongly suggest that P(i) uptake in roots of Z. marina is mediated by a high-affinity Na(+)-dependent transport system. Both NO(3)(-) and P(i) transport systems exploit the steep inwardly directed electrochemical potential gradient for Na(+), considering the low cytoplasmic Na(+) activity (10.7+/-3.3 mM Na(+)) and the high external Na(+) concentration (500 mM Na(+)).  相似文献   

3.
The reduction of nitrate to nitrite in the bacterial periplasm occurs in the 90 kDa NapA subunit of the periplasmic nitrate reductase (NAP) system. Most Shewanella genomes contain two nap operons: napEDABC and napDAGHB, which is an unusual feature of this genus. Two native, monomeric, 90 kDa nitrate reductase active proteins were resolved by hydrophobic interaction chromatography from aerobic cultures of Shewanella gelidimarina replete with reduced nitrogen compounds. The 90 kDa protein obtained in higher yield was characterized as NapA by electronic absorption and electron paramagnetic resonance spectroscopies and was identified by LC/MS/MS and MALDI-TOF/TOF MS as NapA from the napEDABC-type operon. The other 90 kDa protein, which was unstable and produced in low yields, was posited as NapA from the napDAGHB-type operon. Two napA genes have been sequenced from the napEDABC-type and napDAGHB-type operons of S. gelidimarina. Native NAP from S. putrefaciens was resolved as one NapA monomer and one NapAB heterodimer. Two amino acid substitutions in NapA correlated with the isolation of NAP as a NapA monomer or a NapAB heterodimer. The resolution of native, redox-active NapA isoforms in Shewanella provides new insight into the respiratory versatility of this genus, which has implications in bioremediation and the assembly of microbial fuel cells.  相似文献   

4.
The Escherichia coli NapA (periplasmic nitrate reductase) contains a [4Fe-4S] cluster and a Mo-bis-molybdopterin guanine dinucleotide cofactor. The NapA holoenzyme associates with a di-heme c-type cytochrome redox partner (NapB). These proteins have been purified and studied by spectropotentiometry, and the structure of NapA has been determined. In contrast to the well characterized heterodimeric NapAB systems ofalpha-proteobacteria, such as Rhodobacter sphaeroides and Paracoccus pantotrophus, the gamma-proteobacterial E. coli NapA and NapB proteins purify independently and not as a tight heterodimeric complex. This relatively weak interaction is reflected in dissociation constants of 15 and 32 mum determined for oxidized and reduced NapAB complexes, respectively. The surface electrostatic potential of E. coli NapA in the apparent NapB binding region is markedly less polar and anionic than that of the alpha-proteobacterial NapA, which may underlie the weaker binding of NapB. The molybdenum ion coordination sphere of E. coli NapA includes two molybdopterin guanine dinucleotide dithiolenes, a protein-derived cysteinyl ligand and an oxygen atom. The Mo-O bond length is 2.6 A, which is indicative of a water ligand. The potential range over which the Mo(6+) state is reduced to the Mo(5+) state in either NapA (between +100 and -100 mV) or the NapAB complex (-150 to -350 mV) is much lower than that reported for R. sphaeroides NapA (midpoint potential Mo(6+/5+) > +350 mV), and the form of the Mo(5+) EPR signal is quite distinct. In E. coli NapA or NapAB, the Mo(5+) state could not be further reduced to Mo(4+). We then propose a catalytic cycle for E. coli NapA in which nitrate binds to the Mo(5+) ion and where a stable des-oxo Mo(6+) species may participate.  相似文献   

5.
Escherichia coli expresses an inducible flavohemoglobin possessing robust NO dioxygenase activity. At 37 degrees C, the enzyme shows a maximal turnover number (V(max)) of 670 s(-1) and K(m) values for NADH, NO, and O(2) equal to 4.8, 0.28, and approximately 100 microM, respectively. Individual reduction, ligand binding, and NO dioxygenation reactions were examined at 20 degrees C, where V(max) is approximately 94 s(-1). Reduction by NADH occurs in two steps. NADH reduces bound FAD with a rate constant of approximately 15 microM(-1) s(-1), and heme iron is reduced by FADH(2) with a rate constant of 150 s(-1). Dioxygen binds tightly to reduced flavohemoglobin, with association and dissociation rate constants equal to 38 microM(-1) s(-1) and 0.44 s(-1), respectively, and the oxygenated flavohemoglobin dioxygenates NO to form nitrate. NO also binds reversibly to reduced flavohemoglobin in competition with O(2), dissociates slowly, and inhibits NO dioxygenase activity at [NO]/[O(2)] ratios of 1:100. Replacement of the heme pocket B10 tyrosine with phenylalanine increases the O(2) dissociation rate constant approximately 80-fold and reduces NO dioxygenase activity approximately 30-fold, demonstrating the importance of the tyrosine hydroxyl for O(2) affinity and NO scavenging activity. At 37 degrees C, V(max)/K(m)(NO) is 2,400 microM(-1) s(-1), demonstrating that the enzyme is extremely efficient at converting toxic NO into nitrate under physiological conditions.  相似文献   

6.
The periplasmic nitrate reductase, NapA, from Escherichia coli was identified as a 90 kDa molybdoprotein which comigrated during polyacrylamide gel electrophoresis with the di-haem c-type cytochrome, NapB. The DNA sequence of the 5' end of the napA gene and the N-terminal amino acid sequences of both NapA and NapB were determined. The 36 residue leader peptide for NapA includes the double-arginine motif typical of proteins to which complex redox cofactors are attached in the cytoplasm prior to targeting to the periplasm. The pre-NapA leader sequence is both unexpectedly long and, unless two successive proteolysis steps are involved, is cleaved at the unprecedented sequence G-Q-Q-. Nap activity was suppressed during growth in the presence of tungstate and was absent from a mutant unable to synthesise the molybdopterin cofactor.  相似文献   

7.
The cytochrome P450 CYP79B1 from Sinapis alba has been heterologously expressed in Escherichia coli and shown to catalyze the conversion of tryptophan to indole-3-acetaldoxime. Three expression constructs were made, one expressing the native protein and two expressing proteins with different N-terminal modifications. The native construct gave the highest yield as estimated by enzymatic activity per liter of culture. Spheroplasts of E. coli expressing CYP79B1 were reconstituted with the Arabidopsis thaliana NADPH:cytochrome P450 reductase ATR1 heterologously expressed in E. coli to obtain enzymatic activity. This indicates that the E. coli electron-donating system, flavodoxin/flavodoxin reductase, does not support CYP79B1 activity. Recombinant CYP79B1 has a K(m) for tryptophan of 29+/-2 microM and a V(max) of 36.5+/-0.7nmolh(-1)(mlculture)(-1). The identity at the amino acid level of CYP79B1 is, respectively, 93 and 84% to CYP79B2 and CYP79B3 from A. thaliana, and 96% to CYP79B5 (Accession No. AF453287) from Brassica napus. The CYP79B subfamily of cytochromes P450 is likely to constitute a group of orthologous genes in the biosynthesis of indole glucosinolates.  相似文献   

8.
Assimilatory NADH:nitrate reductase (EC 1.6.6.1), a complex molybdenum-, cytochrome b(557)- and FAD-containing protein, catalyzes the regulated and rate-limiting step in the utilization of inorganic nitrogen by higher plants. To facilitate structure/function studies of the individual molybdenum center, we have developed bacterial expression systems for the heterologous production of the 541 residue amino-terminal, molybdenum center-containing domain of spinach nitrate reductase either as a six-histidine-tagged variant or as a glutathione-S-transferase-tagged fusion protein. Expression of the his-tagged molybdenum domain in Escherichia coli BL21(DE3) cells under anaerobic conditions yielded a 55-kDa domain with a specific activity of 1.5 micromol NO(3)(-) consumed/min/nmol enzyme and with a K(mapp)(NO(3)(-)) of 8 mciroM. In contrast, expression of the molybdenum domain as a GST-tagged fusion protein in E. coli TP1000(MobA(-) strain) cells under aerobic conditions yielded an 85-kDa fusion protein with a specific activity of 10.8 micromol NO(3)(-) consumed/min/nmol enzyme and with a K(mapp)(NO(3)(-)) of 12 microM. Fluorescence analysis indicated that both forms of the molybdenum domain contained the cofactor, MPT, although the MPT content was higher in the GST-fusion domain. Inductively coupled plasma mass spectrometric analysis of both the his-tagged and GST-fusion protein domain samples indicated Mo/protein ratios of 0.44 and 0.93, respectively, confirming a very high level of Mo incorporation in the GST-fusion protein. Expression of the GST-fusion protein in TP1000 cells in the presence of elevated tungsten concentrations resulted in an 85-kDa fusion protein that contained MPT but which was devoid of nitrate-reducing activity. Partial reduction of the molybdenum domain resulted in the generation of an axial Mo(V) EPR species with g values of 1.9952, 1.9693, and 1.9665, respectively, and exhibiting superhyperfine coupling to a single exchangeable proton, analogous to that previously observed for the native enzyme. In contrast, the tungsten-substituted MPT-containing domain yielded a W(V) EPR species with g values of 1.9560, 1.9474, and 1.9271, respectively, with unresolved superhyperfine interaction. NADH:nitrate reductase activity could be reconstituted using the GST-molybdenum domain fusion protein in the presence of the recombinant forms of the spinach nitrate reductase' flavin- and heme-containing domains.  相似文献   

9.
The rumen bacterium Wolinella succinogenes grows by respiratory nitrate ammonification with formate as electron donor. Whereas the enzymology and coupling mechanism of nitrite respiration is well known, nitrate reduction to nitrite has not yet been examined. We report here that intact cells and cell fractions catalyse nitrate and chlorate reduction by reduced viologen dyes with high specific activities. A gene cluster encoding components of a putative periplasmic nitrate reductase system (napA, G, H, B, F, L, D) was sequenced. The napA gene was inactivated by inserting a kanamycin resistance gene cassette. The resulting mutant did not grow by nitrate respiration and did not reduce nitrate during growth by fumarate respiration, in contrast to the wild type. An antigen was detected in wild-type cells using an antiserum raised against the periplasmic nitrate reductase (NapA) from Paracoccus pantotrophus. This antigen was absent in the W. succinogenes napA mutant. It is concluded that the periplasmic nitrate reductase NapA is the only respiratory nitrate reductase in W. succinogenes, although a second nitrate-reducing enzyme is apparently induced in the napA mutant. The nap cluster of W. succinogenes lacks a napC gene whose product is thought to function in quinol oxidation and electron transfer to NapA in other bacteria. The W. succinogenes genome encodes two members of the NapC/NirT family, NrfH and FccC. Characterization of corresponding deletion mutants indicates that neither of these two proteins is required for nitrate respiration. A mutant lacking the genes encoding respiratory nitrite reductase (nrfHA) had wild-type properties with respect to nitrate respiration. A model of the electron transport chain of nitrate respiration is proposed in which one or more of the napF, G, H and L gene products mediate electron transport from menaquinol to the periplasmic NapAB complex. Inspection of the W. succinogenes genome sequence suggests that ammonia formation from nitrate is catalysed exclusively by periplasmic respiratory enzymes.  相似文献   

10.
The yields of nitrate and nitrite from decomposition of peroxynitrite in phosphate buffer at 37 degrees C were determined in the pH range 1-14. The NO(2)(-)/NO(3)(-) yields showed a stepwise variation with pH, with inflection points at approximately pH 3.1, 5.8, 6.8, 8.0, and 11.9. Nitrite formation increased strongly above pH 7 at the expense of nitrate, but above pH 12 nitrate again became the major product (80% at pH 14). At this pH, the Arrhenius parameters were E(a)=24.1+/-0.2kcal mol(-1) and A=(4.9+/-1.3)x10(12)s(-1). The yields of NO(2)(-), NO(3)(-), and O(2) measured at pH 5.8, 7.4, and 8.5 as a function of the initial peroxynitrite concentration (50-1000 microM) were linear only at pH 5.8. In the presence of carbon dioxide, oxygen production at pH 7.5 and pH 10 was found to be linear on the CO(2) concentration. The experimental observations were satisfactorily reproduced by kinetic simulations including principal component analyses. These data strongly suggest that the chemistry of peroxynitrite is exclusively mediated by z.rad;NO(2) and HO(z.rad;) radicals in the absence, and by z.rad;NO(2) and CO(3)(z.rad;-) radicals in the presence of CO(2).  相似文献   

11.
The periplasmic nitrate reductase (Nap) is wide-spread in proteobacteria. NapA, the nitrate reductase catalytic subunit, contains a Mo-bisMGD cofactor and one [4Fe-4S] cluster. The nap gene clusters in many bacteria, including Rhodobacter sphaeroides DSM158, contain an napF gene, disruption of which drastically decreases both in vitro and in vivo nitrate reductase activities. In spite its importance in the Nap system, NapF has never been characterized biochemically, and its role remains unknown. The NapF protein has four polycysteine clusters that suggest that it is an iron-sulfur-containing protein. In the present study, a His(6)-tagged NapF protein was overproduced in Escherichia coli and purified anaerobically. The purified NapF protein was used to obtain polyclonal antibodies raised in rabbit, and cellular fractionation of R. sphaeroides followed by immunoprobing with anti-NapF antibodies revealed that the native NapF protein is located in the cytoplasm. This contrasts with the periplasmic location of the mature NapA. However, NapA could not be detected in an isogenic napF(-) strain of R. sphaeroides. The His(6)-tagged NapF protein displayed spectral properties indicative of Fe-S clusters, but these features were rapidly lost, suggesting cluster lability. However, reconstitution of the Fe-S centers into the apo-NapF protein was achieved in the presence of Azotobacter vinelandii cysteine desulfurase (NifS), and this allowed the recovery of nitrate reductase activity in NapA protein that had previously been treated with 2,2'-dipyridyl to remove the [4Fe-4S] cluster. This activity was not recovered in the absence of NapF. Taking into account the cytoplasmic localization of NapF, the presence of labile Fe-S clusters in the protein, the napF(-) strain phenotype, and the NapF-dependent reactivation of the 2,2'-dipyridyl-treated NapA, we propose a role for NapF in assembling the [4Fe-4S] center of the catalytic subunit NapA.  相似文献   

12.
Lansdon EB  Fisher AJ  Segel IH 《Biochemistry》2004,43(14):4356-4365
Recombinant human 3'-phosphoadenosine 5'-phosphosulfate (PAPS) synthetase, isoform 1 (brain), was purified to near-homogeneity from an Escherichia coli expression system and kinetically characterized. The native enzyme, a dimer with each 71 kDa subunit containing an adenosine triphosphate (ATP) sulfurylase and an adenosine 5'-phosphosulfate (APS) kinase domain, catalyzes the overall formation of PAPS from ATP and inorganic sulfate. The protein is active as isolated, but activity is enhanced by treatment with dithiothreitol. APS kinase activity displayed the characteristic substrate inhibition by APS (K(I) of 47.9 microM at saturating MgATP). The maximum attainable activity of 0.12 micromol min(-1) (mg of protein)(-1) was observed at an APS concentration ([APS](opt)) of 15 microM. The theoretical K(m) for APS (at saturating MgATP) and the K(m) for MgATP (at [APS](opt)) were 4.2 microM and 0.14 mM, respectively. At likely cellular levels of MgATP (2.5 mM) and sulfate (0.4 mM), the overall endogenous rate of PAPS formation under optimum assay conditions was 0.09 micromol min(-1) (mg of protein)(-1). Upon addition of pure Penicillium chrysogenum APS kinase in excess, the overall rate increased to 0.47 micromol min(-1) (mg of protein)(-1). The kinetic constants of the ATP sulfurylase domain were as follows: V(max,f) = 0.77 micromol min(-1) (mg of protein)(-1), K(mA(MgATP)) = 0.15 mM, K(ia(MgATP)) = 1 mM, K(mB(sulfate)) = 0.16 mM, V(max,r) = 18.7 micromol min(-1) (mg of protein)(-1), K(mQ(APS)) = 4.8 microM, K(iq(APS)) = 18 nM, and K(mP(PPi)) = 34.6 microM. The (a) imbalance between ATP sulfurylase and APS kinase activities, (b) accumulation of APS in solution during the overall reaction, (c) rate acceleration provided by exogenous APS kinase, and (d) availability of both active sites to exogenous APS all argue against APS channeling. Molybdate, selenate, chromate ("chromium VI"), arsenate, tungstate, chlorate, and perchlorate bind to the ATP sulfurylase domain, with the first five serving as alternative substrates that promote the decomposition of ATP to AMP and PP(i). Selenate, chromate, and arsenate produce transient APX intermediates that are sufficiently long-lived to be captured and 3'-phosphorylated by APS kinase. (The putative PAPX products decompose to adenosine 3',5'-diphosphate and the original oxyanion.) Chlorate and perchlorate form dead-end E.MgATP.oxyanion complexes. Phenylalanine, reported to be an inhibitor of brain ATP sulfurylase, was without effect on PAPS synthetase isoform 1.  相似文献   

13.
Nitric oxide (NO) is generated in biological systems primarily via the activity of NO synthases and nitrate and nitrite reductases. Here we show that Salmonella enterica serovar Typhimurium (S. typhimurium) grown anaerobically with nitrate is capable of generating polarographically detectable NO after nitrite (NO(2)(-)) addition. NO accumulation is sensitive to the NO scavenger 2-(4-carboxyphenyl)-4,4,5,5-tetramethylimidazoline-1-oxyl-3-oxide. Neither an fnr mutant nor an fnr hmp double mutant produces NO, indicating the involvement in NO evolution from NO(2)(-) of protein(s) positively regulated by FNR. Contrary to previous findings in Escherichia coli, we demonstrate that neither the periplasmic nitrite reductase (NrfA) nor the cytoplasmic nitrite reductase (NirB) is involved in NO production in S. typhimurium. However, mutant cells lacking the membrane-bound nitrate reductase, NarGHI, and membranes derived from these cells are unable to produce NO, demonstrating that, in wild-type S. typhimurium, this enzyme is responsible for NO production. Membrane terminal oxidases cannot account for the NO levels measured. The nitrate reductase inhibitor, azide, abrogates NO evolution by Salmonella, and production of NO occurs only in the absence from the assays of nitrate; both features reveal a marked similarity between the NO-generating activities of this bacterium and plants. Unlike the situation in E. coli, an S. typhimurium hmp mutant produces NO both aerobically and anaerobically. Under aerobic conditions, when a functional flavohemoglobin is present, no NO is detectable. We propose a homeostatic mechanism in S. typhimurium, in which NO produced from NO(2)(-) by nitrate reductase derepresses Hmp expression (via FNR and NsrR) and NorV expression (via NorR) and thus limits NO toxicity.  相似文献   

14.
Escherichia coli grew anaerobically on a minimal medium with glycerol as the carbon and energy source and dimethyl sulfoxide (DMSO) as the terminal electron acceptor. DMSO reductase activity, measured with an artificial electron donor (reduced benzyl viologen), was preferentially associated with the membrane fraction (77 +/- 10% total cellular activity). A Km for DMSO reduction of 170 +/- 60 microM was determined for the membrane-bound activity. Methyl viologen, reduced flavin mononucleotide, and reduced flavin adenine dinucleotide also served as electron donors for DMSO reduction. Methionine sulfoxide, a DMSO analog, could substitute for DMSO in both the growth medium and in the benzyl viologen assay. DMSO reductase activity was present in cells grown anaerobically on DMSO but was repressed by the presence of nitrate or by aerobic growth. Anaerobic growth on DMSO coinduced nitrate, fumarate, and and trimethylamine-N-oxide reductase activities. The requirement of a molybdenum cofactor for DMSO reduction was suggested by the inhibition of growth and a 60% reduction in DMSO reductase activity in the presence of 10 mM sodium tungstate. Furthermore, chlorate-resistant mutants chlA, chlB, chlE, and chlG were unable to grow anaerobically on DMSO. DMSO reduction appears to be under the control of the fnr gene.  相似文献   

15.
Disruption of the Aspergillus nidulans high-affinity nitrate transporter genes (nrtA and nrtB) prevents growth on nitrate but not nitrite. We identified a distinct nitrite transporter (K(m)=4.2+/-1 microM, V(max)=168+/-21 nmolmg(-1)DW(-1)h(-1)), designated NitA. Disruption of nrtA, nrtB and nitA blocked growth on nitrite, despite low rates of nitrite depletion we ascribe to passive nitrous acid permeation. Growth of the single mutant nitA16 on nitrite was wild-type, suggesting that NrtA and/or NrtB transports nitrite as well as nitrate. Indeed, NrtA and NrtB transport nitrite at higher rates than NitA; K(m) and V(max) values were 16+/-4 microM and 808+/-67 nmolmg(-1)DW(-1)h(-1) (NrtA) and 11+/-1 microM and 979+/-17 nmolmg(-1)DW(-1)h(-1) (NrtB). We suggest that NrtA is a nitrate/nitrite transporter, NrtB absorbs nitrite in preference to nitrate and NitA is exclusively a nitrite transporter.  相似文献   

16.
A soluble F(1)-ATPase was isolated from the mitochondria of crayfish (Orconectes virilis) gill tissue. The maximal mitochondrial disruption rate (95%) was obtained by sonicating for 4 min at pH 8.6. A 15-fold purification was estimated. The properties for both soluble and membrane-bound enzyme were studied. Both enzyme forms were stable at 4 to -70 degrees C when kept in 20% glycerol. Soluble F(1)-ATPase was more stable at room temperature than membrane-bound enzyme. It displayed a narrower pH profile (pK(1) =6.58, pK(2)=7.68) and more acid pH optimum (7.13) than membrane-bound enzyme (pK(1)=6.42, pK(2)=8.55, optimum pH 7.49). The anion-stimulated activities were in the order HCO(3)(-)>SO(4)(2-)>Cl(-). The apparent K(a) values for soluble enzyme were 11.4, 11.2, and 10.9 mM, respectively, but the K(a) of HCO(3)(-) for membrane-bound enzyme (14.9 mM) was higher than for soluble enzyme. Oligomycin and DCCD inhibited membrane-bound F(1)-ATPase with I(50) of 18.6 ng/ml and 2.2 microM, respectively, but were ineffective in inhibiting soluble enzyme. Both enzyme forms shared identical sensitivity to DIDS (I(50)=12.5 microM) and vanadate (I(50)=9.0 mM). Soluble ATPase was significantly more sensitive to pCMB (I(50)=0.15 microM) and NO(3)(-) (I(50)=28.6 mM) than membrane-bound enzyme (I(50)=1.04 microM pCMB and 81.5 mM NO(3)(-)). In addition, soluble F(1)-ATPase was slightly more sensitive to azide (I(50)=91.8 microM) and NBD-Cl (I(50)=9.18 microM) than membrane-bound enzyme (I(50)=111.6 microM azide and 12.88 microM NBD-Cl). These data suggest a conformational change transmission between F(0) and F(1) sectors and slight conformational differences between soluble F(1) and membrane-bound F(1). In addition, an unmodified F(0) stabilizes F(1) and decreases F(1) sensitivities to inhibitors and modulators.  相似文献   

17.
Disruption of the Aspergillus nidulans high-affinity nitrate transporter genes (nrtA and nrtB) prevents growth on nitrate but not nitrite. We identified a distinct nitrite transporter (K(m)=4.2+/-1 microM, V(max)=168+/-21 nmolmg(-1)DW(-1)h(-1)), designated NitA. Disruption of nrtA, nrtB and nitA blocked growth on nitrite, despite low rates of nitrite depletion we ascribe to passive nitrous acid permeation. Growth of the single mutant nitA16 on nitrite was wild-type, suggesting that NrtA and/or NrtB transports nitrite as well as nitrate. Indeed, NrtA and NrtB transport nitrite at higher rates than NitA; K(m) and V(max) values were 16+/-4 microM and 808+/-67 nmolmg(-1)DW(-1)h(-1) (NrtA) and 11+/-1 microM and 979+/-17 nmolmg(-1)DW(-1)h(-1) (NrtB). We suggest that NrtA is a nitrate/nitrite transporter, NrtB absorbs nitrite in preference to nitrate and NitA is exclusively a nitrite transporter.  相似文献   

18.
The reaction of nitric oxide (NO) with fast and reduced cytochrome bo(3)(cyt bo(3)) from Escherichia coli has been investigated. The stoichiometry of NO binding to cyt bo(3) was determined using an NO electrode in the [NO] range 1-14 microM. Under reducing conditions, the initial decrease in [NO] following the addition of cyt bo(3) corresponded to binding of 1 NO molecule per cyt bo(3) functional unit. After this "rapid" NO binding phase, there was a slow, but significant rate of NO consumption ( approximately 0.3molNOmol bo(3)(-1)min(-1)), indicating that cyt bo(3) possesses a low level of NO reductase activity. The binding of NO to fast pulsed enzyme was also investigated. The results show that in the [NO] range used (1-14 microM) both fast and pulsed oxidised cyt bo(3) bind NO with a stoichiometry of 1:1 with an observed dissociation constant of K(d)=5.6+/-0.6 microM and that NO binding was inhibited by the presence of Cl(-). The binding of nitrite to the binuclear centre causes spectral changes similar to those observed upon NO binding to fast cyt bo(3). These results are discussed in relation to the model proposed by Wilson and co-workers [FEBS Lett. 414 (1997) 281] where the binding of NO to Cu(B)(II) results in the formation of the nitrosonium (Cu(B)(I)-NO(+)) complex. NO(+) then reacts with OH(-), a Cu(B) ligand, to form nitrite, which can bind at the binuclear centre. This work suggests for the first time that the binding of NO to oxidised cyt bo(3) does result in the reduction of Cu(B).  相似文献   

19.
We developed a gas chromatography-mass spectrometry (GC-MS) assay to measure the activity of malonyl-coenzyme A (CoA) decarboxylase (MCD) in crude tissue homogenates. Liver extracts are incubated with [U-(13)C(3)]malonyl-CoA to form [U-(13)C(2)]acetyl-CoA by the action of MCD. The reaction mixture contains 2 mM ADP to prevent the hydrolysis of [1,2-(13)C(2)]acetyl-CoA by acetyl-CoA hydrolase present in the extracts. Newly formed [U-(13)C(2)]acetyl-CoA and internal standard of [(2)H(3),1-(13)C]acetyl-CoA are analyzed as thiophenol derivatives by GC-MS. This assay was applied to a study of the kinetics of MCD in rat liver. Using the Lineweaver-Burke plot of MCD kinetics, K(m) of 202microM and V(max) of 3.3micromol min(-1) (g liver)(-1) were calculated. The liver MCD activities (micromol min(-1) g(-1)+/-SD) in three groups of rats with different nutritional statuses-fed, 1-day fasted, and 2-day fasted-were 1.80+/-0.41, 2.59+/-0.37 (P<0.05), and 3.07+/-0.70 (P<0.05), respectively. We report a practical, nonradioactive, sensitive assay of MCD in crude tissue extract.  相似文献   

20.
Bacterial iodate (IO(3)(-)) reduction is poorly understood largely due to the limited number of available isolates as well as the paucity of information about key enzymes involved in the reaction. In this study, an iodate-reducing bacterium, designated strain SCT, was newly isolated from marine sediment slurry. SCT is phylogenetically closely related to the denitrifying bacterium Pseudomonas stutzeri and reduced 200 microM iodate to iodide (I(-)) within 12 h in an anaerobic culture containing 10 mM nitrate. The strain did not reduce iodate under the aerobic conditions. An anaerobic washed cell suspension of SCT reduced iodate when the cells were pregrown anaerobically with 10 mM nitrate and 200 microM iodate. However, cells pregrown without iodate did not reduce it. The cells in the former category showed methyl viologen-dependent iodate reductase activity (0.31 U mg(-1)), which was located predominantly in the periplasmic space. Furthermore, SCT was capable of anaerobic growth with 3 mM iodate as the sole electron acceptor, and the cells showed enhanced activity with respect to iodate reductase (2.46 U mg(-1)). These results suggest that SCT is a dissimilatory iodate-reducing bacterium and that its iodate reductase is induced by iodate under anaerobic growth conditions.  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