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 375 毫秒
1.
(1) Under anaerobic conditions the respiratory chain in cells of Paracoccus denitrificans, from late exponential cultures grown anaerobically with nitrate as electron acceptor and succinate as carbon source, has been shown to reduce added nitrate via nitrite and nitrous oxide to nitrogen without any accumulation of these intermediates. (2) Addition of nitrous oxide to cells reducing nitrate strongly inhibited the latter reaction. The inhibition was reversed by preventing electron flow to nitrous oxide with either antimycin or acetylene. Electron flow to nitrous oxide thus resembles electron flow to oxygen in its inhibitory effect on nitrate reduction. In contrast, addition of nitrite to an anaerobic suspension of cells reducing nitrate resulted in a stimulation of nitrate reductase activity. Usually, addition of nitrite also partially overcame the inhibitory effect of nitrous oxide on nitrate reduction. The reason why added nitrous oxide, but not nitrite, inhibits nitrate reduction is suggested to be related to the higher reductase activity of the cells for nitrous oxide compared with nitrite. Explanations for the unexpected stimulation of nitrate reduction by nitrite in the presence or absence of added nitrous oxide are considered. (3) Nitrous oxide reductase was shown to be a periplasmic protein that competed with nitrite reductase for electrons from reduced cytochrome c. Added nitrous oxide strongly inhibited the reduction of added nitrite. (4) Nitrite reductase activity of cells was strongly inhibited by oxygen in the presence of physiological reductants, but nitrite reduction did occur in the presence of oxygen when isoascorbate plus N,N,N′,N′-tetramethyl-p-phenylenediamine was the reductant. It is concluded that competition for available electrons by two oxidases, cytochrome aa3 and cytochrome o, severely restricted electron flow to the nitrite reductase (cytochrome cd). For this reason it is unlikely that the oxidase activity of this cytochrome is ever functional in cells. (5) The mechanism by which electron flow to oxygen or nitrous oxide inhibits nitrate reduction in cells has been investigated. It is argued that relatively small changes in the extent of reduction of ubiquinone, or of another component of the respiratory chain with similar redox potential, critically determine the capacity for reducing nitrate. The argument is based on: (i) the response of an anthroyloxystearic acid fluorescent probe that is sensitive to changes in the oxidation state of ubiquinone; (ii) consideration of the total rates of electron flow through ubiquinone both in the presence of oxygen and in the presence of nitrate under anaerobic conditions; (iii) use of relative extents of oxidation of b-type cytochromes as an indicator of ubiquinone redox state, especially the finding that b-type cytochrome of the antimycin-sensitive part of the respiratory chain is more oxidised in the presence of added nitrous oxide, which inhibits nitrate reduction, than in the presence of added nitrite which does not inhibit. Arguments against b- or c-type cytochromes themselves controlling nitrate reduction are given. (6) In principle, control on nitrate reduction could be exerted either upon electron flow or upon the movement of nitrate to the active site of its reductase. The observations that inverted membrane vesicles and detergent-treated cells reduced nitrate and oxygen simultaneously at a range of total rates of electron flow are taken to support the latter mechanism. The failure of an additional reductant, durohydroquinone, to activate nitrate reduction under aerobic conditions in the presence of succinate is also evidence that it is not an inadequate supply of electrons that prevents the functioning of nitrate reductase under aerobic conditions. (7) In inverted membrane vesicles the division of electron flow between nitrate and oxygen is determined by a competition mechanism, in contrast to cells. This change in behaviour upon converting cells to vesicles cannot be attributed to loss of cytochrome c, and therefore of oxidase activity, from the vesicles because a similar change in behaviour was seen with vesicles prepared from cells of a cytochrome c-deficient mutant.  相似文献   

2.
Thauera selenatis grows anaerobically with selenate, nitrate or nitrite as the terminal electron acceptor; use of selenite as an electron acceptor does not support growth. When grown with selenate, the product was selenite; very little of the selenite was further reduced to elemental selenium. When grown in the presence of both selenate and nitrate both electron acceptors were reduced concomitantly; selenite formed during selenate respiration was further reduced to elemental selenium. Mutants lacking the periplasmic nitrite reductase activity were unable to reduce either nitrite or selenite. Mutants possessing higher activity of nitrite reductase than the wild-type, reduced nitrite and selenite more rapidly than the wild-type. Apparently, the nitrite reductase (or a component of the nitrite respiratory system) is involved in catalyzing the reduction of selenite to elemental selenium while also reducing nitrite. While periplasmic cytochrome C 551 may be a component of the nitrite respiratory system, the level of this cytochrome was essentially the same in mutant and wild-type cells grown under two different growth conditions (i.e. with either selenate or selenate plus nitrate as the terminal electron acceptors). The ability of certain other denitrifying and nitrate respiring bacteria to reduce selenite will also be described.  相似文献   

3.
Reduction of nitrite to nitric oxide by enteric bacteria   总被引:10,自引:0,他引:10  
Seven bacteria representing seven genera of enteric bacteria, in addition to Escherichia coli, were shown to reduce nitrite to NO under anaerobic conditions when the cells were grown as nitrate respirers. NO production was inhibited by nitrate and azide and was self limiting, just as was found to be the case previously with E. coli and its nitrate reductase. Maximum initial rates of NO production were observed at pH 5.5-6.  相似文献   

4.
Preparation of a nitrate reductase lysate of Escherichia coli MC1061 to measure nitrate and nitrite in biologic fluids is described. To obtain the crude bacterial lysate containing nitrate reductase activity, E. coli MC1061 was subjected to 16-20 freeze-thawing cycles, from -70 to 60 degrees C, until nitrite reductase activity was < or = 25%. Nitrate reductase activity was detected mainly in the crude preparation. To validate the nitrate reduction procedure, standard nitrate solutions (1.6-100 microM) were incubated with the nitrate reductase preparation for 3 h at 37 degrees C, and nitrite was estimated by the Griess reaction in a microassay. Nitrate solutions were reduced to nitrite in a range of 60-70%. Importantly, no cofactors were necessary to perform nitrate reduction. The biological samples were first reduced with the nitrate reductase preparation. After centrifugation, samples were deproteinized with either methanol/ether or zinc sulfate and nitrite was quantified. The utility of the nitrate reductase preparation was assessed by nitrate+nitrite determination in serum of animals infected with the protozoan Entamoeba histolytica or the bacteria E. coli and in the supernatant of cultured lipopolysaccharide-stimulated RAW 264.7 mouse macrophages. Our results indicate that the nitrate reductase-containing lysate provides a convenient tool for the reduction of nitrate to determine nitrate+nitrite in biological fluids by spectrophotometric methods.  相似文献   

5.
narK mutants of Escherichia coli produce wild-type levels of nitrate reductase but, unlike the wild-type strain, do not accumulate nitrite when grown anaerobically on a glucose-nitrate medium. Comparison of the rates of nitrate and nitrite metabolism in cultures growing anaerobically on glucose-nitrate medium revealed that a narK mutant reduced nitrate at a rate only slightly slower than that in the NarK+ parental strain. Although the specific activities of nitrate reductase and nitrite reductase were similar in the two strains, the parental strain accumulated nitrite in the medium in almost stoichiometric amounts before it was further reduced, while the narK mutant did not accumulate nitrite in the medium but apparently reduced it as rapidly as it was formed. Under conditions in which nitrite reductase was not produced, the narK mutant excreted the nitrite formed from nitrate into the medium; however, the rate of reduction of nitrate to nitrite was significantly slower than that of the parental strain or that which occurred when nitrite reductase was present. These results demonstrate that E. coli is capable of taking up nitrate and excreting nitrite in the absence of a functional NarK protein; however, in growing cells, a functional NarK promotes a more rapid rate of anaerobic nitrate reduction and the continuous excretion of the nitrite formed. Based on the kinetics of nitrate reduction and of nitrite reduction and excretion in growing cultures and in washed cell suspensions, it is proposed that the narK gene encodes a nitrate/nitrite antiporter which facilitates anaerobic nitrate respiration by coupling the excretion of nitrite to nitrate uptake. The failure of nitrate to suppress the reduction of trimethylamine N-oxide in narK mutants was not due to a change in the level of trimethylamine N-oxide reductase but apparently resulted from a relative decrease in the rate of anaerobic nitrate reduction caused by the loss of the antiporter system.  相似文献   

6.
The metabolically versatile purple bacterium Rhodobacter sphaeroides 2.4.3 is a denitrifier whose genome contains two periplasmic nitrate reductase-encoding gene clusters. This work demonstrates nonredundant physiological roles for these two enzymes. One cluster is expressed aerobically and repressed under low oxygen while the second is maximally expressed under low oxygen. Insertional inactivation of the aerobically expressed nitrate reductase eliminated aerobic nitrate reduction, but cells of this strain could still respire nitrate anaerobically. In contrast, when the anaerobic nitrate reductase was absent, aerobic nitrate reduction was detectable, but anaerobic nitrate reduction was impaired. The aerobic nitrate reductase was expressed but not utilized in liquid culture but was utilized during growth on solid medium. Growth on a variety of carbon sources, with the exception of malate, the most oxidized substrate used, resulted in nitrite production on solid medium. This is consistent with a role for the aerobic nitrate reductase in redox homeostasis. These results show that one of the nitrate reductases is specific for respiration and denitrification while the other likely plays a role in redox homeostasis during aerobic growth.  相似文献   

7.
A denitrifying mutant of Bacillus stearothermophilus NCA 2184, strain 2184-D, was used to explore the development of nitrate respiration in relation to oxygen respiration. Aerobically grown wild-type cultures could acquire the ability to use nitrate as a result of selection of nitrate-respiring mutants by the presence of nitrate and a reduced oxygen tension. Fluctuation analysis has revealed that the frequency of occurrence of the nitrate-respiring mutant is about 7.5 x 10(-8) per bacterium per generation. Nitrate reductase and nitrite reductase appeared to be induced sequentially in strain 2184-D by the addition of nitrate. The formation of both of these enzymes was repressed by oxygen so that cells grown aerobically with nitrate possessed a low basal level of nitrate reducatase and exhibited no denitrification. The rate of synthesis of nitrate reductase increased quickly after addition of nitrate and removal of oxygen. It then declined to a lower steady-state level. Cells grown anaerobically with nitrate retained approximately 30 to 40% of the respiratory activity of aerobically grown cells. Aeration of anaerobically grown cells in the presence of amino acids increased the respiratory activity to normal aerobic levels. This aeration promoted rapid degradation of the existing nitrate reductase with or without the added amino acids.  相似文献   

8.
The denitrifying bacterium Paracoccus denitrificans can grow aerobically or anaerobically using nitrate or nitrite as the sole nitrogen source. The biochemical pathway responsible is expressed from a gene cluster comprising a nitrate/nitrite transporter (NasA), nitrite transporter (NasH), nitrite reductase (NasB), ferredoxin (NasG) and nitrate reductase (NasC). NasB and NasG are essential for growth with nitrate or nitrite as the nitrogen source. NADH serves as the electron donor for nitrate and nitrite reduction, but only NasB has a NADH-oxidizing domain. Nitrate and nitrite reductase activities show the same Km for NADH and can be separated by anion-exchange chromatography, but only fractions containing NasB retain the ability to oxidize NADH. This implies that NasG mediates electron flux from the NADH-oxidizing site in NasB to the sites of nitrate and nitrite reduction in NasC and NasB respectively. Delivery of extracellular nitrate to NasBGC is mediated by NasA, but both NasA and NasH contribute to nitrite uptake. The roles of NasA and NasC can be substituted during anaerobic growth by the biochemically distinct membrane-bound respiratory nitrate reductase (Nar), demonstrating functional overlap. nasG is highly conserved in nitrate/nitrite assimilation gene clusters, which is consistent with a key role for the NasG ferredoxin, as part of a phylogenetically widespread composite nitrate and nitrite reductase system.  相似文献   

9.
Abstract The Gram-positive soil bacterium Bacillus subtilis , generally regarded as an aerobe, grows under strict anaerobic conditions using nitrate as an electron acceptor and should be designated as a facultative anaerobe. Growth experiments demonstrated a lag phase of 24 to 36 hours after the shift from aerobic, to the onset of anaerobic respiratory growth. Anaerobically adapted cells grew without further lag phase after their transfer to fresh anaerobic growth medium. The cells change their morphology from rods to longer filament-like structures when moved from aerobic to anaerobic respiratory growth conditions. Surprisingly, anaerobically grown B. subtilis lost the capacity for sporulation. An investigation of the molecular basis of the switch between aerobic and anaerobic growth was initiated by the cloning of the genes encoding the respiratory nitrate reductase from B. subtilis . Oligonucleotides deduced from conserved amino acid sequence regions of eubacterial respiratory nitrate reductases and related enzymes were used for the isolation of the genes. Four open reading frames with significant homology to the E. coli respiratory nitrate reductase opérons ( narGHIJ, narZYWV ) were isolated and termed narGHJI . A chromosomal knock-out mutation of the B. subtilis nar operon totally abolished nitrate respiration.  相似文献   

10.
Nitric oxide is a key element in host defense against invasive pathogens. The periplasmic cytochrome c nitrite reductase (NrfA) of Escherichia coli catalyzes the respiratory reduction of nitrite, but in vitro studies have shown that it can also reduce nitric oxide. The physiological significance of the latter reaction in vivo has never been assessed. In this study the reduction of nitric oxide by Escherichia coli was measured in strains active or deficient in periplasmic nitrite reduction. Nrf(+) cells, harvested from cultures grown anaerobically, possessed a nitric-oxide reductase activity with physiological electron donation of 60 nmol min(-1) x mg dry wt(-1), and an in vivo turnover number of NrfA of 390 NO* s(-1) was calculated. Nitric-oxide reductase activity could not be detected in Nrf(-) strains. Comparison of the anaerobic growth of Nrf(+) and Nrf(-) strains revealed a higher sensitivity to nitric oxide in the NrfA(-) strains. A higher sensitivity to the nitrosating agent S-nitroso-N-acetyl penicillamine (SNAP) was also observed in agar plate disk-diffusion assays. Oxygen respiration by E. coli was also more sensitive to nitric oxide in the Nrf(-) strains compared with the Nrf(+) parent strain. The results demonstrate that active periplasmic cytochrome c nitrite reductase can confer the capacity for nitric oxide reduction and detoxification on E. coli. Genomic analysis of many pathogenic enteric bacteria reveals the presence of nrf genes. The present study raises the possibility that this reflects an important role for the cytochrome c nitrite reductase in nitric oxide management in oxygen-limited environments.  相似文献   

11.
Paracoccus pantotrophus grown anaerobically under denitrifying conditions expressed similar levels of the periplasmic nitrate reductase (NAP) when cultured in molybdate- or tungstate-containing media. A native PAGE gel stained for nitrate reductase activity revealed that only NapA from molybdate-grown cells displayed readily detectable nitrate reductase activity. Further kinetic analysis showed that the periplasmic fraction from cells grown on molybdate (3 microM) reduced nitrate at a rate of V(max)=3.41+/-0.16 micromol [NO(3)(-)] min(-1) mg(-1) with an affinity for nitrate of K(m)=0.24+/-0.05 mM and was heat-stable up to 50 degrees C. In contrast, the periplasmic fraction obtained from cells cultured in media supplemented with tungstate (100 microM) reduced nitrate at a much slower rate, with much lower affinity (V(max)=0.05+/-0.002 micromol [NO(3)(-)] min(-1) mg(-1) and K(m)=3.91+/-0.45 mM) and was labile during prolonged incubation at >20 degrees C. Nitrate-dependent growth of Escherichia coli strains expressing only nitrate reductase A was inhibited by sub-mM concentrations of tungstate in the medium. In contrast, a strain expressing only NAP was only partially inhibited by 10 mM tungstate. However, none of the above experimental approaches revealed evidence that tungsten could replace molybdenum at the active site of E. coli NapA. The combined data show that tungsten can function at the active site of some, but not all, molybdoenzymes from mesophilic bacteria.  相似文献   

12.
Proton translocation coupled to the reduction of nitrite was studied in anaerobically grown Escherichia coli. Extrusion of protons occurred by adding nitrite to an anaerobic suspension of wild-type cells. This extrusion was sensitive to a proton conductor, 3,5-di-tert-butyl-4-hydroxybenzylidenemalononitrile (SF6847) or carbonylcyanide-p-trifluoromethoxyphenylhydrazone. Dicyclohexylcarbodiimide, an inhibitor of H+-ATPase, prevented the proton extrusion linked to nitrite reduction, whereas this reagent had no effect on respiratory nitrate reduction to nitrite. Proton extrusion was undetectable when nitrite was added to a suspension of mutant cells defective in H+-ATPase. These results indicate that the proton extrusion associated with nitrite reduction to ammonia is not by redox pumps but by H+-ATPase. From the results obtained by the measurement of proton extrusion in nitrite reductase-deficient mutants, NADH-nitrite reductase system is suggested to involve the proton extrusion in whole cells of E. coli.  相似文献   

13.
14.
Escherichia coli grown anaerobically on nitrate exhibited the same transport barrier to reduction of chlorate, relative to nitrate, as that exhibited by Paracoccus denitrificans. This establishes that the nitrate binding site of nitrate reductase (EC 1.7.99.4) in E. coli must also lie on the cell side of the nitrate transporter which is associated with the plasma membrane. Because nitrate reductase is membrane bound, the nitrate binding site is thus located on the inner aspect of the membrane. Nitrate pulse studies on E. coli in the absence of valinomycin showed a small transient alkalinization (leads to H+/NO3- congruent to --0.07) which did not occur with oxygen pulses. By analogy with P. denitrificans, the alkaline transient is interpreted to arise from proton-linked nitrate uptake which is closely followed by nitrite efflux. The result is consistent with internal reduction of nitrate, whereas external reduction would be expected to give leads to H+/NO3-ratios approaching --2.  相似文献   

15.
Two polytopic membrane proteins, NarK and NarU, are assumed to transport nitrite out of the Escherichia coli cytoplasm, but how nitrate enters enteric bacteria is unknown. We report the construction and use of four isogenic strains that lack nitrate reductase Z and the periplasmic nitrate reductase, but express all combinations of narK and narU. The active site of the only functional nitrate reductase, nitrate reductase A, is located in the cytoplasm, so nitrate reduction by these four strains is totally dependent upon a mechanism for importing nitrate. These strains were exploited to determine the roles of NarK and NarU in both nitrate and nitrite transport. Single mutants that lack either NarK or NarU were competent for nitrate-dependent anaerobic growth on a non-fermentable carbon source, glycerol. They transported and reduced nitrate almost as rapidly as the parental strain. In contrast, the narK-narU double mutant was defective in nitrate-dependent growth unless nitrate transport was facilitated by the nitrate ionophore, reduced benzyl viologen (BV). It was also unable to catalyse nitrate reduction in the presence of physiological electron donors. Synthesis of active nitrate reductase A and the cytoplasmic, NADH-dependent nitrite reductase were unaffected by the narK and narU mutations. The rate of nitrite reduction catalysed by the cytoplasmic, NADH-dependent nitrite reductase by the double mutant was almost as rapid as that of the NarK+-NarU+ strain, indicating that there is a mechanism for nitrite uptake by E. coli that is in-dependent of either NarK or NarU. The nir operon encodes a soluble, cytoplasmic nitrite reductase that catalyses NADH-dependent reduction of nitrite to ammonia. One additional component that contributes to nitrite uptake was shown to be NirC, the hydrophobic product of the third gene of the nir operon, which is predicted to be a polytopic membrane protein with six membrane-spanning helices. Deletion of both NarK and NirC decreased nitrite uptake and reduction to a basal rate that was fully restored by a single chromosomal copy of either narK or nirC. A multicopy plasmid encoding NarU complemented a narK mutation for nitrite excretion, but not for nitrite uptake. We conclude that, in contrast to NirC, which transports only nitrite, NarK and NarU provide alternative mechanisms for both nitrate and nitrite transport. However, NarU might selectively promote nitrite ex-cretion, not nitrite uptake.  相似文献   

16.
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   

17.
The denitrifying ability of thirteen strains of Rhizobium meliloti was tested. Most of the strains were able to reduce nitrate to nitrous oxide or dinitrogen. However, they failed to use nitrate as electron acceptor for ATP generation or growth at low oxygen tensions. Under micro-aerobic conditions, free-living cells of R. meliloti 102-F-51 strain exhibited a constitutive nitrate reductase activity independent of the presence of nitrate. On the other hand, nitrite reductase activity was dependent not only on low levels of oxygen but also on the presence of a high nitrate concentration in the medium. Denitrification activity proceeded immediately once a threshold level of nitrite was accumulated in the medium or in cells incubated with 1mM nitrite. However, a lag period was required when cells were incubated with nitrate.  相似文献   

18.
In anaerobically grown Paracoccus denitrificans the dissimilatory nitrate reductase is linked to the respiratory chain at the level of cytochromes b. Electron transport to nitrite and nitrous oxide involves c-type cytochromes. During electron transport from NADH to nitrate one phosphorylation site is passed, whereas two sites are passed during electron transport from NADH to oxygen, nitrite and nitrous oxide. The presentation of a respiratory chain as a linear array of electron carriers gives a misleading picture of the efficiency of energy conservation since the location of the reductases is not taken into account. For the reduction of nitrite and nitrous oxide, protons are utilized from the periplasmic space, whereas for the reduction of oxygen and nitrate, protons are utilized from the cytoplasmic side of the inner membrane. Evidence for two transport systems for nitrate was obtained. One is driven by the proton motive force; this system is used to initiate nitrate reduction. The second system is a nitrate-nitrite antiport system. A scheme for proton translocation and electron transport to nitrate, nitrite, nitrous oxide and oxygen is presented. The number of charges translocated across the membrane during flow of two electrons from NADH is the same for all nitrogenous oxides and is 67-71% of that during electron transfer to oxygen via cytochrome o. These findings are in accordance with growth yield studies. YMAX electron values determined in chemostat cultures for growth with various substrates and hydrogen acceptors are proportional to the number of charges translocated to these hydrogen acceptors during electron transport.  相似文献   

19.
The reduction of (per)chlorate and nitrate in (per)chlorate-reducing bacteria shows similarities and differences. (Per)chlorate reductase and nitrate reductase both belong to the type?II DMSO family of enzymes and have a common bis(molybdopterin guanine dinucleotide)molybdenum cofactor. There are two types of dissimilatory nitrate reductases. With respect to their localization, (per)chlorate reductase is more similar to the dissimilatory periplasmic nitrate reductase. However, the periplasmic, unlike the membrane-bound, respiratory nitrate reductase, is not able to use chlorate. Structurally, (per)chlorate reductase is more similar to respiratory nitrate reductase, since these reductases have analogous subunits encoded by analogous genes. Both periplasmic (per)chlorate reductase and membrane-bound nitrate reductase activities are induced under anoxic conditions in the presence of (per)chlorate and nitrate respectively. During microbial (per)chlorate reduction, molecular oxygen is generated. This is not the case for nitrate reduction, although an atypical reaction in nitrite reduction linked to oxygen formation has been described recently. Microbial oxygen production during reduction of oxyanions may enhance biodegradation of pollutants under anoxic conditions.  相似文献   

20.
A chlorate-resistant mutant B25 of Arabidopsis thaliana (L.) Heinh. was isolated, which has very little or no in vitro nitrate reductase activity and grows poorly on a substrate with nitrate as the sole nitrogen source. The mutation of B25 ( rgn ) is monogenic and recessive, tightly linked to the marker gene an on chromosome 1. Nitrate induces cytochrome- c reductase activity in the mutant but to a lower level than in the wildtype. After sucrose gradient centrifugation the greatest part of the cytochrome- c reductase from induced wildtype is found as 8s type whereas cytochrome- c reductase from B25 under the same conditions is found as 4s type. Nitrate reductase is found at the 8s position. It is suggested that B25 has lost the ability to assemble two 4s subunits showing cytochrome- c reductase activity and a Mo-bearing co-factor into the functional nitrate reductase. Nitrate rather than nitrite is the inducing agent for nitrite reductase, since in B25 nitrite reductase is even more rapidly induced than in the wildtype after addition of nitrate. Both the wildtype and B25 contain a nitrate reductase inhibiting factor when grown on ammonium. This inhibiting factor is a small protein, possibly similar to the nitrate reductase inactivating enzyme reported for other plants.  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