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.
The respiratory nitrate reductase from Paracoccus denitrificans has been purified in the non-ionic detergent Nonidet P-40. The enzyme comprises three polypeptides, alpha, beta and gamma with estimated relative molecular masses of 127 000, 61 000 and 21 000. Duroquinol or reduced-viologen compounds acted as the reducing substrates. The nitrate reductase contained a b-type cytochrome that was reduced by duroquinol and oxidised by nitrate. A preparation of the enzyme that lacked both detectable b-type cytochrome and the gamma subunit was obtained from a trailing peak of nitrate reductase activity collected from a gel filtration column. Absence of the gamma subunit correlated with failure to use duroquinol as reductant; activity with reduced viologens was retained. It is concluded that in the plasma membrane of P. denitrificans the gamma subunit catalyses electron transfer to the alpha and beta subunits of nitrate reductase from ubiquinol which acts as a branch point in the respiratory chain. A new assay was introduced for both nitrate and quinol-nitrate oxidoreductase activity. Diaphorase was used to couple the oxidation of NADH to the production of duroquinol which acted as electron donor to nitrate reductase. Under anaerobic conditions absorbance changes at 340 nm were sensitive to nitrate concentrations in the low micromolar range. This coupled assay was used to determine that the purified enzyme had Km(NO-3) of 13 microM and a Km of 470 microM for ClO-3, an alternative substrate. With viologen substrates Km(NO-3) of 283 microM and Km(ClO-3) of 470 microM were determined; the enzymes possessed a considerably higher Vmax with either NO-3 or ClO-3 than was found when duroquinol was substrate. Azide was a competitive inhibitor of nitrate reduction in either assay system (Ki = 0.55 microM) but 2-n-heptyl-4-hydroxyquinoline N-oxide was effective only with the complete three-subunit enzyme and duroquinol as substrate, consistent with a site of action for this inhibitor on the b-type cytochrome. The low Km for nitrate observed in the duriquinol assay is comparable with the apparent Km(NO-3) recently reported for intact cells of P. denitrificans [Parsonage, D., Greenfield, A. J. & Ferguson, S. J. (1985) Biochim. Biophys. Acta 807, 81-95]. This similarity is discussed in terms of a possible requirement for a nitrate transport system. The nitrate reductase system from P. denitrificans is compared with that from Escherichia coli.  相似文献   

2.
The NAD(P)H-dependent nitrate reductase system in Clostridium perfringens was reconstituted with rubredoxin (Rd), nitrate reductase (NaR), and an unadsorbed fraction, on a DEAE-cellulose column, of the extract (designated as fraction A), under nitrogen gas. Ferredoxin in place of Rd was not effective as an electron carrier in this reconstituted system. NAD(P)H-dependent nitrate reducing activity was also obtained by replacing fraction A with ferredoxin-NADP+ reductase from spinach. We propose the following scheme for the electron transfer in this NAD(P)H dependent nitrate reduction system. NAD(P)H----NAD(P)H-Rd reductase----Rd----NaR----NO3-.  相似文献   

3.
Assimilatory NADH:nitrate reductase (EC 1.6.6.1), a complex Mo-pterin-, cytochrome b(557)-, and FAD-containing protein, catalyzes the regulated and rate-limiting step in the utilization of inorganic nitrogen by higher plants. A codon-optimized gene has been synthesized for expression of the central cytochrome b(557)-containing fragment, corresponding to residues A542-E658, of spinach assimilatory nitrate reductase. While expression of the full-length synthetic gene in Escherichia coli did not result in significant heme domain production, expression of a Y647* truncated form resulted in substantial heme domain production as evidenced by the generation of "pink" cells. The histidine-tagged heme domain was purified to homogeneity using a combination of NTA-agarose and size-exclusion FPLC, resulting in a single protein band following SDS-PAGE analysis with a molecular mass of approximately 13 kDa. MALDI-TOF mass spectrometry yielded an m/z ratio of 12,435 and confirmed the presence of the heme prosthetic group (m/z=622) while cofactor analysis indicated a 1:1 heme to protein stoichiometry. The oxidized heme domain exhibited spectroscopic properties typical of a b-type cytochrome with a visible Soret maximum at 413 nm together with epr g-values of 2.98, 2.26, and 1.49, consistent with low-spin bis-histidyl coordination. Oxidation-reduction titrations of the heme domain indicated a standard midpoint potential (E(o)') of -118 mV. The isolated heme domain formed a 1:1 complex with cytochrome c with a K(A) of 7 microM (micro=0.007) and reconstituted NADH:cytochrome c reductase activity in the presence of a recombinant form of the spinach nitrate reductase flavin domain, yielding a k(cat) of 1.4 s(-1) and a K(m app) for cytochrome c of 9 microM. These results indicate the efficient expression of a recombinant form of the heme domain of spinach nitrate reductase that retained the spectroscopic and thermodynamic properties characteristic of the corresponding domain in the native spinach enzyme.  相似文献   

4.
The microsomal enzyme 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase catalyzes the rate-limiting step in the cholesterogenic pathway and was proposed to be composed in situ of 2 noncovalently linked subunits (Edwards, P.A., Kempner, E.S., Lan, S.-F., and Erickson, S.K. (1985) J. Biol. Chem. 260, 10278-10282). In the present report, the activities and kinetic properties of HMG-CoA reductase in microsomes isolated from livers of rats fed on diets supplemented with either ground Amberlite XAD-2 ("X"), cholestyramine/mevinolin ("CM"), or unsupplemented, normal rat chow ("N"), were compared. The specific activities of HMG-CoA reductase in X and CM microsomes were, respectively, 5- and 83-fold higher than that of N microsomes. In NADPH-dependent kinetics of HMG-CoA reductase activated with 4.5 mM GSH, the concentration of NADPH required for half-maximal velocity (S0.5) was 209 +/- 23, 76 +/- 23, and 40 +/- 4 microM for the N, X, and CM microsomes, respectively. While reductase from X microsomes displays cooperative kinetics toward NADPH (Hill coefficient (nH) = 1.97 +/- 0.07), the enzyme from CM microsomes does not (nH = 1.04 +/- 0.07). Similarly to HMG-CoA reductase from CM microsomes, the freeze-thaw solubilized enzyme ("SOL") displays no cooperativity toward NADPH and its Km for this substrate is 34 microM. At 4.5 mM GSH, HMG-CoA reductase from X, CM, and SOL preparations has a similar Km value for [DL]-HMG-CoA, ranging between 13-16 microM, while reductase from N microsomes had a higher Km value (42 microM) for this substrate. No cooperativity towards HMG-CoA was observed in any of the tested enzyme preparations. Immunoblotting analyses of the different preparations demonstrated that the observed altered kinetics of HMG-CoA reductase in the microsomes is not due to preferential proteolytic cleavage of the native 97-100 kDa subunit of the enzyme to the noncooperative 50-55 kDa species. Moreover, it was found that the ratio enzymatic activity/immunoreactivity of the reductase increased in the order N less than X less than CM approximately equal to SOL, indicating that the activity per reductase molecule increases with the induction of the enzyme. These results are compatible with a model suggesting that dietary induction of hepatic HMG-CoA reductase may change the state of functional aggregation of its subunits.  相似文献   

5.
The biosynthetic replacement of Met residues by selenomethionine (SeMet) facilitates the determination of three-dimensional structure by multiwavelength anomalous diffraction (Yang, W., Hendrickson, W. A., Crouch, R.J., and Satow, Y. (1990) Science 249, 1398-1405). In an effort to examine any biochemical effects due to the replacement of Met residues by SeMet, we chose to compare the kinetic and binding properties of selenomethionyl dihydrofolate reductase with those of the wt enzyme. There are 5 Met residues in Escherichia coli dihydrofolate reductase with 2 located in the Met-20 loop, which is a sequence of residues forming a lid over the active site. Utilizing plasmid pWT8, which affords 10-15% soluble protein as E. coli dihydrofolate reductase, we readily isolated both the SeMet and wt enzymes from E. coli DL41 utilizing a novel purification protocol. Both enzymes exhibited essentially the same kinetic and binding properties, including specific activities (45 mumol/min/mg), Km (7,8-dihydrofolate = 0.39 microM; NADPH = 2.0 microM), kcat (13.5/s), and 1:1 noncovalent inhibitory binding ratios with methotrexate. The inhibitory effects of divalent and monovalent cations on activity were also assessed, with the SeMet-containing enzyme exhibiting a uniformly greater sensitivity than the wt enzyme. We conclude that the biochemical properties of dihydrofolate reductase are virtually unperturbed by SeMet inclusion. Analysis of SeMet dihydrofolate reductase by 77Se nuclear magnetic resonance spectroscopy revealed five distinct resonances, thus indicating the potential value of this technique in employing selenium as a nonperturbing NMR probe of protein structure and function.  相似文献   

6.
In Escherichia coli, aerobiosis inhibits the synthesis of enzymes for anaerobic respiration (e.g., nitrate reductase and fumarate reductase) and for fermentation (e.g., formate-hydrogen lyase). Anaerobically, nitrate induces nitrate reductase synthesis and inhibits the formation of both fumarate reductase and formate-hydrogen lyase. Previous work has shown that narL+ is required for the effects of nitrate on synthesis of both nitrate reductase and fumarate reductase. Another gene, narK (whose function is unknown), has no observable effect on formation of these enzymes. We report here our studies on the role of nar genes in fumarate reductase and formate-hydrogen lyase gene expression. We observed that insertions in narX (also of unknown function) significantly relieved nitrate inhibition of fumarate reductase gene expression. This phenotype was distinct from that of narL insertions, which abolished this nitrate effect under certain growth conditions. In contrast, insertion mutations in narK and narGHJI (the structural genes for the nitrate reductase enzyme complex) significantly relieved nitrate inhibition of formate-hydrogen lyase gene expression. Insertions in narL had a lesser effect, and insertions in narX had no effect. We conclude that nitrate affects formate-hydrogen lyase synthesis by a pathway distinct from that for nitrate reductase and fumarate reductase.  相似文献   

7.
8.
The effect of six compounds isolated from rhizome of Anemone raddeana on the superoxide generation in human neutrophils was investigated. The six compounds examined were 3-acetyloleanolic acid (AOA), oleanolic acid (OA), eleutheroside K (EK), oleanolic acid-3-O-alpha-L-rhamnopyranosyl-(1 --> 2)-[beta-D-glucopyranosyl-(1 --> 4)]-alpha-L-arabinopyranoside (Rd10), raddeanoside 12 (Rd12) and raddeanoside 13 (Rd13). AOA, OA, Rd12 and Rd13 suppressed the superoxide generation induced by N-formyl-methionyl-leucyl-phenylalanine (fMLP) in a concentration-dependent manner. EK and Rd10 significantly enhanced the fMLP-induced superoxide generation in a specific narrow range of low concentration (0.5-0.75 microM), while these compounds more efficiently suppressed the superoxide generation than the other four compounds in other concentrations. In the case of superoxide generation induced by phorbol 12-myristate 13-acetate (PMA), Rd12, OA, EK and Rd10 dose-dependently suppressed the superoxide generation but AOA and Rd13 gave no effect. Arachidonic acid-induced superoxide generation was suppressed by EK, Rd10, Rd12 and Rd13, but was weakly enhanced by AOA and OA. Rd12 dose-dependently inhibited fMLP-induced tyrosyl phosphorylation of 123.0, 79.4, 60.3, 56.2 and 50.1 kDa proteins in human neutrophil. On the other hand, RD10 and EK enhanced the tyrosyl phosphorylation of these proteins in a low concentration range. These phenomena were parallel to the suppression of the fMLP-induced superoxide generations.  相似文献   

9.
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   

10.
The ferredoxin-dependent nitrate reductase from the cyanobacterium Synechococcus sp. PCC 7942 has been shown to form a high-affinity complex with ferredoxin at low ionic strength. This complex, detected by changes in both the absorbance and circular dichroism (CD) spectra, did not form at high ionic strength. When reduced ferredoxin served as the electron donor for the reduction of nitrate to nitrite, the activity of the enzyme declined markedly as the ionic strength increased. In contrast, the activity of the enzyme with reduced methyl viologen (a non-physiological electron donor) was independent of ionic strength. These results suggest that an electrostatically stabilized complex between Synechococcus nitrate reductase and ferredoxin plays an important role in the mechanism of nitrate reduction catalyzed by this enzyme. Treatment of Synechococcus nitrate reductase with either an arginine-modifying reagent or a lysine-modifying reagent inhibited the ferredoxin-dependent activity of the enzyme but did not affect the methyl viologen-dependent activity. Treatment with these reagents also resulted in a large decrease in the affinity of the enzyme for ferredoxin. Formation of a nitrate reductase complex with ferredoxin prior to treatment with either reagent protected the enzyme against loss of ferredoxin-dependent activity. These results suggest that lysine and arginine residues are present at the ferredoxin-binding site of Synechococcus nitrate reductase. Results of experiments using site-specific, charge reversal variants of the ferredoxin from the cyanobacterium Anabaena sp. PCC 7119 as an electron donor to nitrate reductase were consistent with a role for negatively charged residues on ferredoxin in the interaction with Synechococcus nitrate reductase.  相似文献   

11.
The function of a highly mobile loop in Escherichia coli dihydrofolate reductase was studied by constructing a mutant (DL1) using cassette mutagenesis that had four residues deleted in the middle section of the loop (Met16-Ala19) and a glycine inserted to seal the gap. This part of the loop involves residues 16-20 and is disordered in the X-ray crystal structures of the apoprotein and the NADP+ binary complex but forms a hairpin turn that folds over the nicotinamide moiety of NADP+ and the pteridine moiety of folate in the ternary complex [Bystroff, C., & Kraut, J. (1991) Biochemistry 30, 2227-2239]. The steady-state and pre-steady-state kinetics and two-dimensional 1H NMR spectra were analyzed and compared to the wild-type protein. The kinetics on the DL1 mutant enzyme show that the KM value for NADPH (5.3 microM), the KM for dihydrofolate (2 microM), the rate constant for the release of the product tetrahydrofolate (10.3 s-1), and the intrinsic pKa value (6.2) are similar to those exhibited by the wild-type enzyme. However, the hydride-transfer rate declines markedly from the wild-type value of 950 s-1 to 1.7 s-1 for the DL1 mutant and when taken with data for substrate binding indicates that the loop contributes to substrate flux by a factor of 3.5 x 10(4). Thus, the mobility of loop I may provide a mechanism of recruiting hydrophobic residues which can properly align the nicotinamide and pteridine rings for the hydride-transfer process (a form of transition-state stabilization).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Thioltransferase in human red blood cells: purification and properties   总被引:3,自引:0,他引:3  
Thioltransferase activity was identified and the enzyme purified to apparent homogeneity from human red blood cells. Activity was measured as glutathione-dependent reduction of the prototype substrate hydroxyethyl disulfide; formation of oxidized glutathione (GSSG) was coupled to NADPH oxidation by GSSG reductase (1 unit of activity = 1 mumol/min of NADPH oxidized). The thioltransferase-GSH-GSSG reductase system was shown also to catalyze the regeneration of hemoglobin from the mixed disulfide hemoglobin-S-S-glutathione (HbSSG) and to reactivate the metabolic control enzyme phosphofructokinase (PFK) after oxidation of its sulfhydryl groups. On a relative concentration basis, thioltransferase was about 1200 times more efficient than dithiothreitol in reactivation of phosphofructokinase; e.g., 500 microM DTT was required to effect the same extent of reactivation as that of 0.4 microM TTase. The GSH plus GSSG reductase system without thioltransferase was ineffective for reduction of HbSSG or reactivation of PFK. The average amount of thioltransferase in intact erythrocytes was calculated to be 4.6 units/g of Hb at 25 degrees C. This level of activity is about the same as those of other enzymes that participate in sulfhydryl maintenance in red blood cells, such as GSSG reductase and glucose-6-phosphate dehydrogenase. These results suggest a physiological role for the thioltransferase in erythrocyte sulfhydryl homeostasis. Certain properties of the human erythrocyte thioltransferase resemble those of other mammalian thioltransferase and glutaredoxin enzymes. Thus, the human erythrocyte enzyme, purified about 28,000-fold to apparent homogeneity, is a single polypeptide with a molecular weight of 11,300. Its N-terminus is blocked, it is heat stable, and it contains four cysteine residues per protein molecule. However, the human erythrocyte thioltransferase is a distinct protein based on its amino acid composition. For example, it contains no methionine residues; whereas the related mammalian enzymes described to date have at least one internal methionine residue in their largely homologous sequences.  相似文献   

13.
14.
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   

15.
The primary structure of prostaglandin (PG) F synthetase from bovine lung shows 62% similarity with that of human liver aldehyde reductase (EC 1.1.1.2) (Watanabe, K., Fujii, Y., Nakayama, K., Ohkubo, H., Kuramitsu, S., Kagamiyama, H., Nakanishi, S., and Hayaishi, O. (1988) Proc. Natl. Acad. Sci. U.S.A. 85, 11-15). We therefore purified human liver aldehyde reductase to homogeneity and compared the immunological and catalytic properties of aldehyde reductase and PGF synthetase. Although both enzymes belong to a group of aldoketoreductases and their molecular weights are essentially identical, aldehyde reductase had no cross-reactivity to anti-PGF synthetase antiserum. Furthermore, there was a difference in the substrate specificity for reduction of PGs between the two enzymes. Aldehyde reductase catalyzed the reduction of PGJ2, delta 12-PGJ2, PGH2, or PGA2, but not that of PGB2, PGD2, or PGE2, whereas PGF synthetase reduced PGD2. The optimum pH, Km value for PGH2, and the turnover number were 6.5, 100 microM, and 3.1 min-1, respectively. The PGH2 9,11-endoperoxide reductase activity of aldehyde reductase was not affected in the presence of a substrate such as p-nitrobenzaldehyde, DL-glyceraldehyde, or 9,10-phenanthrenequinone, suggesting that PGH2 9,11-endoperoxide and other substrates are reduced at different active site(s). The reaction product formed from PGH2 by this enzyme was identified as PGF2 alpha by gas chromatography/mass spectrometry. These results suggest that aldehyde reductase is not exactly identical to PGF synthetase in terms of its immunological property and substrate specificity for PGs, but that this enzyme is also involved in the direct conversion of PGH2 to PGF2 alpha similar to PGF synthetase.  相似文献   

16.
Desulfovibrio desulfuricans ATCC 27774 is a sulfate reducer that can adapt to nitrate respiration, inducing the enzymes required to utilize this alternative metabolic pathway. Nitrite reductase from this organism has been previously isolated and characterized, but no information was available on the enzyme involved in the reduction of nitrate. This is the first report of purification to homogeneity of a nitrate reductase from a sulfate reducing organism, thus completing the enzymatic system required to convert nitrate (through nitrite) to ammonia. D. desulfuricans nitrate reductase is a monomeric (circa 70 kDa) periplasmic enzyme with a specific activity of 5.4 K(m) for nitrate was estimated to be 20 microM. EPR signals due to one [4Fe-4S] cluster and Mo(V) were identified in dithionite reduced samples and in the presence of nitrate.  相似文献   

17.
We have compared the characteristics of nitrate uptake by Aphanothece halophytica grown under non-stress and salt-stress conditions. Both cell types showed essentially similar patterns of nitrate uptake toward ammonium, nitrite, and DL-glyceraldehyde. Although the affinities of nitrate to non-stress cells and salt-stress cells were not significantly different, i.e., Ks = 416 and 450 microM, respectively, the V(max) value for non-stress cells was about twofold of that for salt-stress cells (9.1 vs 5.3 micromol min(-1) mg(-1) Chl). Nitrate uptake by A. halophytica was found to be dependent on Na+. Ammonium inhibited nitrate uptake, and the presence of methionine sulfoximine could not release the inhibition by ammonium. Nitrite appeared to competitively inhibit nitrate uptake with a K(i) value of 84 microM. Both chloride and phosphate anions did not affect nitrate uptake. DL-Glyceraldehyde, an inhibitor of CO2 fixation, caused a reduction in the uptake of nitrate.  相似文献   

18.
Trypanothione reductase of Trypanosoma cruzi is a key enzyme in the antioxidant metabolism of the parasite. Here we report on the enzymic and pharmacological properties of trypanothione reductase using glutathionylspermidine disulfide as a substrate. 1. Both pH optimum (7.5) and the ionic strength optimum (at 30 mM) are unusually narrow for this enzyme. 40 mM Hepes, 1 mM EDTA, pH 7.5 was chosen as a standard assay buffer because in this system the kcat/Km ratio had the highest values for both natural substrates, glutathionylspermidine disulfide (2.65 x 10(6) M-1 s-1) and trypanothione disulfide (4.63 x 10(6) M-1 s-1). 2. Using the standardized assay, trypanothione reductase and the phylogenetically related host enzyme, human glutathione reductase, were studied as targets of inhibitors. Both enzymes, in their NADPH-reduced forms, were irreversibly modified by the cytostatic agent, 1,3-bis(2-chloroethyl)-1-nitrosourea (BCNU). Nifurtimox, the drug used in the treatment of Chagas' disease, is a stronger inhibitor of glutathione reductase (Ki = 40 microM) than of trypanothione reductase (IC50 = 200 microM). 3. Of the newly synthesized trypanocidal compounds [Henderson, G. B., Ulrich, P., Fairlamb, A. H., Rosenberg, I., Pereira, M., Sela, M. & Cerami, A. (1988) Proc. Natl Acad. Sci., 85, 5374-5378] a nitrofuran derivative, 2-(5-nitro-2-furanylmethylidene)-N,N'-[1,4-piperazinediylbis (1,3-propanediyl)]bishydrazinecarboximidamide tetrahydrobromide, was found to be a better inhibitor for trypanothione reductase (Ki = 0.5 microM) than for glutathione reductase (IC50 = 10 microM). A naphthoquinone derivative, 2,3-bis[3-(2-amidinohydrazono)-butyl]-1,4-naphthoquinone dihydrochloride, turned out to be both an inhibitor (IC50 = 1 microM) and an NADPH-oxidation-inducing substrate (Km = 14 microM). This effect was not observed with human glutathione reductase. Such compounds which lead to oxidative stress by more than one mechanism in the parasite are promising starting points for drug design based on the three-dimensional structures of glutathione and trypanothione reductases.  相似文献   

19.
The distribution of nitrate reductase (NR, EC 1.6.6.1.) in the leaves of single-stem tomato plants ( Lycopersicon esculentum Mill., cv. Vandenbergs Moneydor) was studied using an in vitro test. The activity decreased from young to old leaves. However, a low value (NR minimum) occurred in some leaves below the apex, usually in the almost completely expanded leaves, provided that the plants received sufficient nitrate to induce optimum NR activity in all the leaves. When insufficient nitrate was available there was NR in the young leaves only. The observed NR minimum coincided with a low value for soluble carbohydrates and amino acids. Since there was no extra export of labelled carbon from the leaves with the NR minimum, it is suggested that in the almost completely expanded leaves carbohydrates produced by photosynthesis are mainly used for the production of polysaccharides for new cell walls. Consequently, less are left for the production of keto acids, which can act as acceptors for reduced nitrogen. Therefore, less amino acids are produced, and this may result in a lowered protein synthesis, including a lowered synthesis of nitrate reductase.  相似文献   

20.
Abstract Nitrate reductase was purified from and characterized in a bloom-forming unicellular calcifying alga, Emiliania huxleyi (Haptophyceae). The molecular masses of the native form and the subunit were 514 and 85 kDa, respectively, showing that the enzyme is a hexamer composed of 6 homologous subunits. The K m values for NADH and NO3− were 40 μM and 104 μM, respectively. Activity of the reduction of nitrate was very high with reduced methylviologen and NADH, but no activity was observed with NADPH or reduced flavin mononucleotide; oxidation of NADH was very high with cytochrome c but did not occur with ferricyanide. These results indicate that Emiliania nitrate reductase is NADH-specific (EC 1.6.6.1), and that among algae and plants its subunit structure and kinetic properties are unique.  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