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.
Harper JE 《Plant physiology》1981,68(6):1488-1493
Studies were conducted to quantitate the evolution of nitrogen oxides (NO(x)) from soybean [Glycine max (L.) Merr.] leaves during in vivo nitrate reductase (NR) assays with aerobic and anaerobic gas purging. Anaerobic gas purging (N2 and argon) consistently resulted in greater NO(x) evolution than did aerobic gas purging (air and O2). The evolution of NO(x) was dependent on gas flow rate and on NO2 formation in the assay medium; although a threshold level of NO2 appeared to exist beyond which the rate of NO(x) evolution did not increase further.  相似文献   

2.
The intracellular distribution of NADPH- and NADH-dependentduroquinone reductase (NAD (P)H-DQR) from etiolated zucchinihypocotyls (Cucurbita pepo L.) was investigated. About 80% ofthis enzyme is in the supernatant fraction and is probably cytosolic.Particulate NAD (P)H-DQR was largely (42%) found in associationwith the plasma membrane and was strongly stimulated by TX100.Another 33% of NAD (P)H-DQR was associated with mitochondria,and minor fractions with the endoplasmic reticulum (8%) andother particles. All these fractions were little or not stimulatedby TX100. The distribution of detergent-activated NAD (P)H-DQRis thus distinct from microsomal NADH- and NADPH-CCR. The plasma membrane was purified from microsomal fractions bymetrizamide plus sucrose density gradient centrifugation orby PEG/dextran phase partitioning. Both types of particle preparationspeaked at a density (d) of 1.165 g cm–3 in sucrose gradientsand contained substantial TX100-sensitive NADH-DQR, TX100-stimulatedNAD (P)H-DQR, together with traces of NADH-CCR and trapped ‘soluble’enzyme (MDH, NADP-malic enzyme) activities. In isopycnic gradientsof unfractionated microsomes, however, trapped enzymes peakedat d 1.155 whereas NAD (P)H-DQR peaked at d 1.165 and GSII atd 1.170, probably revealing plasma membrane heterogeneity. Furtherevidence of heterogeneity was provided by fractionation of plasmamembrane vesicles on dextran step-gradients. Most of the trapped MDH was released to the supernatant by sonicationor treatment with 0.0125% TX100. Under these conditions mostof the NAD (P)H-DQR sedimented with the membranes. It is concludedthat NAD (P)H-DQR is bound to the inside of plasma membranevesicles, but a fraction (7 to 31%) may be ‘soluble’and sequestered within the vesicle lumen. Part of the detergent-sensitiveNADH-DQR may be externally bound and accessible to non-permeatingsubstrates. Key words: Cucurbita, NAD (P)H-quinone reductase, plasma membrane  相似文献   

3.
The cotyledons of soybean begin to develop photosynthetic capacity shortly after emergence. The cotyledons develop nitrate reductase (NR) activity in parallel with an increase in chlorophyll and a decrease in protein. In crude extracts of 5- to 8-day-old cotyledons, NR activity is greatest with NADH as electron donor. In extracts of older cotyledons, NR activity is greatest with NADPH. Blue-Sepharose was used to purify and separate the NR activities into two fractions. When the blue-Sepharose was eluted with NADPH, NR activity was obtained which was most active with NADPH as electron donor. Assays of the NADPH-eluted NR with different concentrations of nitrate revealed that the highest activity was obtained in 80 millimolar KNO3. Thus, this fraction has properties similar to the low nitrate affinity NAD(P)H:NR of soybean leaves. When 5- to 8-day-old cotyledons were extracted and purified, further elution of the blue-Sepharose with KNO3, subsequent to the NADPH elution, yielded an NR fraction most active with NADH. Assays of this fraction with different nitrate concentrations revealed that this NR had a higher nitrate affinity and was similar to the NADH:NR of soybean leaves. The KNO3-eluted NR fraction which was purified from the extracts of 9- to 14-day-old cotyledons, was most active with NADPH. The analysis of these fractions prepared from the extracts of older cotyledons indicated that residual NAD(P)H:NR contaminated the NADH:NR. Despite this complication, the pattern of development of the purified NR fractions was consistent with the changes observed in the crude extract NR activities. It was concluded that NADH:NR was most active in young cotyledons and that as the cotyledons aged the NAD(P)H:NR became more active.  相似文献   

4.
5.
Ferredoxin-dependent nitrite reductase of spinach has been further characterized and the relationship between this enzyme and methyl viologen-dependent nitrite reductase studied.

Purified ferredoxin nitrite reductase, having a molecular weight of 86,000, showed 2.5 times higher ferredoxin-dependent activity than methyl viologen-linked activity. Besides 4 mol of labile sulfide the enzyme contained about 2 mol of siroheme per mol. When dithionite, methyl viologen and nitrite were added, ESR signals of a heme nitrosyl complex at g = 2.14, 2.07 and 2.02 were observed. Moreover, hyperfine splitting of the signal due to 14N nuclear spin was also observed at 2.033, 2.023 and 2.013. The sole addition of hydroxylamine to the ferric enzyme also caused the same but much less intense signals with the hyperfine splitting.

On treatment of the ferredoxin nitrite reductase (native enzyme) with DEAE-Sephadex A-50 chromatography, a modified nitrite reductase having a molecular weight of 61,000 and a protein fraction having an apparent molecular weight of 24,000 were separated. The modified enzyme contained about one mol of siroheme and 4 mol of labile sulfide per mol and showed essentially the same heme ESR signals as the native enzyme. Contrary to the native enzyme, this modified enzyme accepted electrons more efficiently from methyl viologen than ferredoxin and the reduction of nitrite to ammonia catalyzed by the modified enzyme was not stoichiometric. The observed nitrite to ammonia ratio was 1 to less than 0.6. Cyanide at concentrations between 0.02 to 0.2 mm inhibited the activity of the native enzyme almost completely but the modified enzyme was inhibited only partially.

From the results obtained, it is suggested that the native ferredoxin-linked nitrite reductase consists of two components (or subunits) and removal of the light component results in formation of a modified enzyme with increased relative affinity to methyl viologen.  相似文献   

6.
The nitrate reductase activity of 5-day-old whole corn roots was isolated using phosphate buffer. The relatively stable nitrate reductase extract can be separated into three fractions using affinity chromatography on blue-Sepharose. The first fraction, eluted with NADPH, reduces nearly equal amounts of nitrate with either NADPH or NADH. A subsequent elution with NADH yields a nitrate reductase which is more active with NADH as electron donor. Further elution with salt gives a nitrate reductase fraction which is active with both NADH and NADPH, but is more active with NADH. All three nitrate reductase fractions have pH optima of 7.5 and Stokes radii of about 6.0 nanometers. The NADPH-eluted enzyme has a nitrate Km of 0.3 millimolar in the presence of NADPH, whereas the NADH-eluted enzyme has a nitrate Km of 0.07 millimolar in the presence of NADH. The NADPH-eluted fraction appears to be similar to the NAD(P)H:nitrate reductase isolated from corn scutellum and the NADH-eluted fraction is similar to the NADH:nitrate reductases isolated from corn leaf and scutellum. The salt-eluted fraction appears to be a mixture of NAD(P)H: and NADH:nitrate reductases.  相似文献   

7.
The reduction of plastoquinone by NADPH was detected as an increasein the dark level of Chi fluorescence in osmotically rupturedchloroplasts of spinach. This activity was observed only whenthe chloroplasts were ruptured in a medium containing a highconcentration of MgCl2. The activity was suppressed by inhibitorsof the respiratory NADH dehydrogenase (NDH) complex in mitochondria,capsaicin and amobarbital, suggesting that the activity wasmediated by chloroplastic NDH complex. Antimycin A, an inhibitorof ferredoxin-quinone reductase (FQR), and the protonophorenigericin also inhibited the increase in Chi fluorescence byNADPH. By contrast, JV-ethylmaleimide (NEM), an inhibitor offerredoxin-NADP+ reductase (FNR), did not suppress the fluorescenceincrease, showing that FNR is not involved in this reaction.When the osmotically ruptured chloroplasts were washed by centrifugation,a further addition of ferredoxin as well as NADPH was requiredfor an increase in fluorescence. This ferredoxin-de-pendentactivity also was suppressed by antimycin A, but only partlyinhibited by capsaicin or amobarbital, suggesting that thisis mediated mainly by FQR. These findings suggest that the NADPH-bindingsubunit of NDH complex is easily dissociated from the thylakoidmembranes during the process of the washing the thylakoids bycentrifugation. 3Present address: Shanghai Institute of Plant Physiology, AcademiaSinica, 300 Fenglin Road, Shanghai 200032, China 5Present address: Department of Biotechnology, Faculty of Engineering,Fukuyama University, 1 Gakuen-cho, Fukuyama, Hiroshima, 729-02Japan  相似文献   

8.
Lipid peroxidation induced by Fe2+ADP in soybean mitochondriais stimulated by pyruvate, malate (in the presence of NAD2+)and by NAD(P)H. This lipid peroxidation is almost completelyinhibited by EDTA indicating that iron is essential. Also salicylhydroxamicacid, a specific inhibitor of the alternate oxidase, is a stronginhibitor of lipid peroxidation but its effect should be relatedto chelation or to a general antioxidant action. Rotenone doesnot show any effect on the malateNAD2+Fe2+ADP-induced lipidperoxidation. From the reported data, it is possible to concludethat, in soybean mitochondria, the peroxidation of unsaturatedlipids can be modulated through a balance between the systemssparking lipid peroxidation, like NAD(P)H Fe2+ADP, and the systemswhich protect against it, i.e. quinones, maintained at the reducedstate by the substrates. (Received April 20, 1987; Accepted July 21, 1987)  相似文献   

9.
We have cloned the nap locus encoding the periplasmic nitrate reductase in Rhodobacter sphaeroides f. sp. denitrificans IL106. A mutant with this enzyme deleted is unable to grow under denitrifying conditions. Biochemical analysis of this mutant shows that in contrast to the wild-type strain, the level of synthesis of the nitrite and N(2)O reductases is not increased by the addition of nitrate. Growth under denitrifying conditions and induction of N oxide reductase synthesis are both restored by the presence of a plasmid containing the genes encoding the nitrate reductase. This demonstrates that R. sphaeroides f. sp. denitrificans IL106 does not possess an efficient membrane-bound nitrate reductase and that nitrate is not the direct inducer for the nitrite and N(2)O reductases in this species. In contrast, we show that nitrite induces the synthesis of the nitrate reductase.  相似文献   

10.
Nitrite was able to strongly inhibit C2H2 reduction by nitrogenase from soybean bacteroids, whereas H2 evolution was unaffected under the same conditions. NO inhibited both C2H2 reduction and H2 evolution; during C2H2 reduction, sensitivity of nitrogenase to NO was higher than to NO2, and the Ki values were, respectively, 0.056 and 0.52 mM. Production of NO resulting from a reduction of NO2 by dithionite in nitrogenase incubations was observed. However, the characteristics of inhibitions and the low level of NO generated by nitrite reduction ruled out the suggestion concerning a direct role of NO to explain the inhibitory effect of NO2 on nitrogenase.  相似文献   

11.
Glutamate toxicity in the N18-RE-105 neuronal cell line results from the inhibition of high-affinity cystine uptake, which leads to a depletion of glutathione and the accumulation of oxidants. Production of superoxides by one-electron oxidation/reduction of quinones is decreased by NAD(P)H:quinone reductase, an enzyme with DT-diaphorase activity. Using glutamate toxicity in N18-RE-105 cells as a model of neuronal oxidative stress, we report that the degree of glutamate toxicity observed is inversely proportional to quinone reductase activity. Induction of quinone reductase activity by treatment with t-butylhydroquinone reduced glutamate toxicity by up to 80%. In contrast, treatment with the quinone reductase inhibitor dicumarol potentiated the toxic effect of glutamate. Measurement of cellular glutathione indicates that increases in its levels are not responsible for the protective effect of t-butylhydroquinone treatment. Because many types of cell death may involve the formation of oxidants, induction of quinone reductase may be a new strategy to combat neurodegenerative disease.  相似文献   

12.
Barley (Hordeum vulgare L.) has two, differentially regulated, nitrate reductase (NR) genes, one encoding the NADH-specific NR (Nar1) and the other encoding the NAD(P)H-bispecific NR (Nar7). Regulation of the two NR genes by nitrate was investigated in wild-type Steptoe and in an NADH-specific NR structural gene mutant (Az12). Gene-specific probes were used to estimate NADH and NAD(P)H NR mRNAs. The kinetics of induction by nitrate were similar for the two NR genes; expression was generally below the limits of detection prior to induction, reached maximum levels after 1 to 2 h of induction in roots and 4 to 8 h of induction in leaves, and then declined to steady-state levels. Derepression of the NAD(P)H NR gene in leaves of the NADH-specific NR gene mutant Az12 did not appear to be associated with changes in nitrate assimilation products or nitrate flux. Nitrate deprivation resulted in rapid decreases in NADH and NAD(P)H NR mRNAs in seedling roots and leaves and equally rapid decreases in the concentration of nitrate in the xylem sap. These results indicate that factors affecting nitrate uptake and transport could have a direct influence on NR expression in barley leaves.  相似文献   

13.
Carbonyl compounds such as alpha-ketoglutarate, pyruvate, oxaloacetate, butyraldehyde, acetaldehyde or acetone react with NAD or NADP to give adducts. Binding studies of adducts to dehydrogenases are performed by means of ultraviolet differential spectroscopy, circular dichroism and spectrofluorimetry. The dehydrogenases show a high degree of binding specificity toward the adducts which contain their specific oxidized substrate and their specific coenzyme. The high selectivity of the dehydrogenases for adducts is evidenced by binding studies of NAD(P)-pyruvate and NAD(P)-alpha-ketoglutarate adducts on glutamate dehydrogenase at pH 7.6 and 8.9. Evidence is presented showing that adducts bind to the active site of the enzymes.  相似文献   

14.
The NAD+/NADH ratio was 12 in whole soybean nodules tissue,but only 2 in bacteroids, as a result of the high concentrationof NADH. By contrast, NADP+/NADPH ratios were less than unityin both nodules and bacteroids, being 0.28 and 0.37, respectively. The adenylate energy charge values in bacteroids and nodules,0.37 and 0.39, respectively, were remarkably low, and were insharp contrast to the normal value of 0.83 in root tissue. (Received July 19, 1988; Accepted March 9, 1989)  相似文献   

15.
The respiratory chain of plant mitochondria differs from that in mammalian mitochondria by containing several rotenone-insensitive NAD(P)H dehydrogenases. Two of these are located on the outer, cytosolic surface of the inner membrane. One is specific for NADH, the other for NADPH. Only the latter is inhibited by diphenyleneiodonium (DPI). Both of these enzymes are normally dependent upon Ca2+ for activity and this constitutes a potentially important mechanism by which the cell can regulate the oxidation of cytosolic NAD(P)H via the concentration of free Ca2+. This and other potential regulatory mechanisms such as the substrate concentration and polyamines are discussed.  相似文献   

16.
The nonenzymatic reduction of nitrosobenzene by NADPH and NADH in aqueous buffer solution at 25°C is described. Both reactants quantitatively convert nitrosobenzene to phenylhydroxylamine. Rate constants for reduction (kr) were determined spectrophotometrically and found to be identical at pH 5.7 and 7.4 and independent of buffer concentration. The values of kNADH (124–149 M?1 sec?1) and kNADPH (131–170 M?1 sec?1) are essentially identical. The reaction is not subject to general catalysis or specific salt effects. The oxidation of phenylhydroxylamine by NAD(P) to nitrosobenzene is only stimulated by a factor of 1.2 over oxidation in its absence (when the ratio of NADP: phenylhydroxylamine was 8:1).  相似文献   

17.
Vanadate-dependent NAD(P)H oxidation, catalyzed by rat liver microsomes and microsomal NADPH-cytochrome P450 reductase (P450 reductase) and NADH-cytochrome b5 reductase (b5 reductase), was investigated. These enzymes and intact microsomes catalyzed NAD(P)H oxidation in the presence of either ortho- or polyvanadate. Antibody to P450 reductase inhibited orthovanadate-dependent NADPH oxidation catalyzed by either purified P450 reductase or rat liver microsomes and had no effect on the rates of NADH oxidation catalyzed by b5 reductase. NADPH-cytochrome P450 reductase catalyzed orthovanadate-dependent NADPH oxidation five times faster than NADH-cytochrome b5 reductase catalyzed NADH oxidation. Orthovanadate-dependent oxidation of either NADPH or NADH, catalyzed by purified reductases or rat liver microsomes, occurred in an anaerobic system, which indicated that superoxide is not an obligate intermediate in this process. Superoxide dismutase (SOD) inhibited orthovanadate, but not polyvanadate-mediated, enzyme-dependent NAD(P)H oxidation. SOD also inhibited when pyridine nucleotide oxidation was conducted anaerobically, suggesting that SOD inhibits vanadate-dependent NAD(P)H oxidation by a mechanism independent of scavenging of O2-.  相似文献   

18.
An aldehyde derivative of riboflavin was covalently attached by reductive alkylation to soluble polycationic supports. The flavopolymers so obtained were stable under operational conditions. The catalytic efficiency towards oxidation of NADH by these flavopolymers was demonstrated, and the kinetic parameters (Km and kcat) revealed an overall catalytic efficiency (kcat/Km) 185-fold greater compared to riboflavin. Various factors affecting the chemical regeneration of NAD+ from NADH such as pH, ionic strength, nature of the buffer etc. were studied. The most interesting result was the highly favourable influence of borate ions which increased the reaction rate by a factor 2-4 compared to the other buffers. The flavopolymers are very effective for in situ recycling of NAD(P)+. With up to 300-fold NADH----NAD+ conversions for the system using yeast alcohol dehydrogenase and up to 1500-fold NADPH----NADP+ regenerations for the system using glucose-6-phosphate dehydrogenase. These flavopolymers are superior to previous chemical recycling systems.  相似文献   

19.
Elevated cholesterol levels promote proinflammatory and prothrombogenic responses in venules and impaired endothelium-dependent arteriolar dilation. Although NAD(P)H oxidase-derived superoxide has been implicated in the altered vascular responses to hypercholesterolemia, it remains unclear whether this oxidative pathway mediates the associated arteriolar dysfunction and platelet adhesion in venules. Platelet and leukocyte adhesion in cremasteric postcapillary venules and arteriolar dilation responses to acetylcholine were monitored in wild-type (WT), Cu,Zn-superoxide dismutase transgenic (SOD-TgN), and NAD(P)H oxidase-knockout (gp91(phox-/-)) mice placed on a normal (ND) or high-cholesterol (HC) diet for 2 weeks. HC elicited increased platelet and leukocyte adhesion in WT mice versus ND. Cytosolic subunits of NAD(P)H oxidase (p47phox and p67phox) were expressed in platelets. This was not altered by hypercholesterolemia; however, platelets and leukocytes from HC mice exhibited elevated generation of reactive oxygen species compared to ND mice. Hypercholesterolemia-induced leukocyte recruitment was attenuated in SOD-TgN-HC and gp91(phox-/-)-HC mice. Recruitment of platelets derived from WT-HC mice in venules of SOD-TgN-HC or gp91(phox-/-)-HC recipients was comparable to ND levels. Adhesion of SOD-TgN-HC platelets paralleled the leukocyte response and was attenuated in SOD-TgN-HC recipients, but not in WT-HC recipients. However, gp91(phox-/-)-HC platelets exhibited low levels of adhesion comparable to those of WT-ND in both hypercholesterolemic gp91(phox-/-) and WT recipients. Arteriolar dysfunction was evident in WT-HC mice, compared to WT-ND. Overexpression of SOD or, to a lesser extent, gp91(phox) deficiency restored arteriolar vasorelaxation responses toward WT-ND levels. These findings reveal a novel role for platelet-associated NAD(P)H oxidase in producing the thrombogenic phenotype in hypercholesterolemia and demonstrate that NAD(P)H oxidase-derived superoxide mediates the HC-induced arteriolar dysfunction.  相似文献   

20.
A dissimilatory nitrite reductase from Haloferax denitrificans was purified to apparent electrophoretic homogeneity. The overall purification was 125-fold with about a 1% recovery of activity. The enzyme, which had a molecular mass of 127 kDa, was composed of a 64-kDa subunit as determined by SDS-PAGE. Although maximum activity occurred in the presence of 4 M NaCl, no activity was lost when the enzyme was incubated in the absence of NaCl. The absorption spectrum had maxima at 462, 594, and 682 nm, which disappeared upon reduction with dithionite. Diethyldithiocarbamate (DDC) was inhibitory, and the addition of copper sulfate to DDC-inhibited enzyme partially restored activity. These results suggest this enzyme is a copper-containing nitrite reductase and is the first such nitrite reductase to be described in an Archeon.  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