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 31 毫秒
1.
At least two hydroxypyruvate reductases (HPRs), differing in specificity for NAD(P)H and (presumably) utilizing glyoxylate as a secondary substrate, were identified by fractionation of crude maize leaf extracts with ammonium sulfate. The NADH-preferring enzyme, which most probably represented peroxisomal HPR, was precipitated by 30 to 45% saturated ammonium sulfate, while most of the NADPH-dependent activity was found in a 45 to 60% precipitate. The HPRs had similar low Kms for hydroxypyruvate (about 0.1 millimolar), regardless of cofactor, while affinities of glyoxylate reductase (GR) reactions for glyoxylate varied widely (Kms of 0.4-12 millimolar) depending on cofactor. At high hydroxypyruvate concentrations, the NADPH-HPR from the 30 to 45% precipitate showed negative cooperativity with respect to this reactant, having a second Km of 6 millimolar. In contrast, NADPH-HPR from the 45 to 60% precipitate was inhibited at high hydroxypyruvate concentrations (K1 of 3 millimolar) and, together with NADPH-GR, had only few, if any, common antigenic determinants with NADH-HPR from the 30 to 45% fraction. Both NADPH-HPR and NADPH-GR activities from the 45 to 60% precipitate were probably carried out by the same enzyme(s), as found by kinetic studies. Following preincubation with NADPH, there was a marked increase (up to sixfold) in activity of NADPH-HPR from either crude or fractionated extracts. Most of this increase could be attributed to an artefact resulting from an interference by endogeneous NADPH-phosphatase, which hydrolyzed NADPH to NADH, the latter being utilized by the NADH-dependent HPR. However, in the presence of 15 millimolar fluoride (phosphatase inhibitor), preincubation with NADPH still resulted in over 60% activation of NADPH-HPR. The NADPH treatment stimulated the Vmax of the reductase but had no effect on its Km for hydroxypyruvate. Enzyme distribution studies revealed that both NADH and NADPH-dependent HPR and GR activities were predominantly localized in the bundle sheath compartment. Rates of NADPH-HPR and NADPH-GR in this tissue (over 100 micromoles per hour per milligram of chlorophyll each) are in the upper range of values reported for leaves of C3 species.  相似文献   

2.
1. Two enzymes that catalyse the reduction of glyoxylate to glycollate have been separated and purified from a species of Pseudomonas. Their molecular weights were estimated as 180000. 2. Reduced nicotinamide nucleotides act as the hydrogen donators for the enzymes. The NADH-linked enzyme is entirely specific for its coenzyme but the NADPH-linked reductase shows some affinity towards NADH. 3. Both enzymes convert hydroxypyruvate into glycerate. 4. The glyoxylate reductases show maximal activity at pH6·0–6·8, are inhibited by keto acids and are strongly dependent on free thiol groups for activity. 5. The Michaelis constants for glyoxylate and hydroxypyruvate were found to be of a high order. 6. The reversibility of the reaction has been demonstrated for both glyoxylate reductases and the equilibrium constants were determined. 7. The reduction of glyoxylate and hydroxypyruvate is not stimulated by anions.  相似文献   

3.
Burke JJ 《Plant physiology》1990,93(2):652-656
The relationship between the thermal dependence of the reappearance of chlorophyll variable fluorescence following illumination and temperature dependence of the apparent Michaelis constant (Km) of NADH hydroxypyruvate reductase for NADH was investigated in cool and warm season plant species. Brancker SF-20 and SF-30 fluorometers were used to evaluate induced fluorescence transients from detached leaves of wheat (Triticum aestivum L. cv TAM-101), cotton (Gossypium hirsutum L. cv Paymaster 145), tomato (Lycopersicon esculentum cv Del Oro), bell pepper (Capsicum annuum L. cv California Wonder), and petunia (Petunia hybrida cv. Red Sail). Following an illumination period at 25°C, the reappearance of variable fluorescence during a dark incubation was determined at 5°C intervals from 15°C to 45°C. Variable fluorescence recovery was normally distributed with the maximum recovery observed at 20°C in wheat, 30°C in cotton, 20°C to 25°C in tomato, 30 to 35°C in bell pepper and 25°C in petunia. Comparison of the thermal response of fluorescence recovery with the temperature sensitivity of the apparent Km of hydroxypyruvate reductase for NADH showed that the range of temperatures providing fluorescence recovery corresponded with those temperatures providing the minimum apparent Km values (viz. the thermal kinetic window).  相似文献   

4.
The use of LaPr 88/29 mutant of barley (Hordeum vulgare), which lacks NADH-preferring hydroxypyruvate reductase (HPR-1), allowed for an unequivocal demonstration of at least two related NADPH-preferring reductases in this species: HPR-2, reactive with both hydroxypyruvate and glyoxylate, and the glyoxylate specific reductase (GR-1). Antibodies against spinach HPR-1 recognized barley HPR-1 and partially reacted with barley HPR-2, but not GR-1, as demonstrated by Western immunoblotting and immunoprecipitation of proteins from crude leaf extracts. The mutant was deficient in HPR-1 protein. In partially purified preparations, the activities of HPR-1, HPR-2, and GR-1 could be differentiated by substrate kinetics and/or inhibition studies. Apparent Km values of HPR-2 for hydroxypyruvate and glyoxylate were 0.7 and 1.1 millimolar, respectively, while the Km of GR-1 for glyoxylate was 0.07 millimolar. The Km values of HPR-1, measured in wild type, for hydroxypyruvate and glyoxylate were 0.12 and 20 millimolar, respectively. Tartronate and P-hydroxypyruvate acted as selective uncompetitive inhibitors of HPR-2 (Ki values of 0.3 and 0.4 millimolar, respectively), while acetohydroxamate selectively inhibited GR-1 activity. Nonspecific contributions of HPR-1 reactions in assays of HPR-2 and GR-1 activities were quantified by a direct comparison of rates in preparations from wild-type and LaPr 88/29 plants. The data are evaluated with respect to previous reports on plant HPR and GR activities and with respect to optimal assay procedures for individual HPR-1, HPR-2, and GR-1 rates in leaf preparations.  相似文献   

5.
Leaf extracts from seven monocotyledonous and dicotyledonous species contained considerable levels of NADPH-dependent glyoxylate- and hydroxypyruvate reductase activities. These activities ranged from 0.02 to 0.22 μmol (mg protein)−1 min−1. For all plants tested, the glyoxylate reductase (GR) activity, assayed with either NADPH or NADH, was sensitive to inhibition by acetohydroxamate, a glycine analogue. Hydroxypyruvate reductase (HPR) activities were unaffected by acetohydroxamate. Differential precipitation of soluble leaf proteins of spinach, pea and barley by ammonium sulfate (0–45% and 45–60% saturation) indicated the presence of at least three distinct reductases, which differed in their specificities for glyoxylate, hydroxypyruvate and NAD(P)H. For all species, the NADH-dependent HPR-activity was almost completely precipitated by low ammonium sulfate concentration (45%), while precipitation of the NADPH-GR, NADH-GR and, to some extent, NADPH-HPR activities required 60% ammonium sulfate. The NADPH-dependent GR and HPR activities had high affinity for glyoxylate and hydroxypyruvate, respectively, as indicated by low apparent Km values of 40–120 μ M . The occurrence of at least three distinct reductases utilizing hydroxypyruvate and/or glyoxylate as substrate was supported by antibody-precipitation studies using antibodies prepared against NADH(NADPH)-HPR, the well-known peroxisomal enzyme that also shows non-specific GR activity. These data are discussed with respect to recent reports on the purification and characterization of NADPH(NADH)-GR, and NADPH (NADH)-HPR, two cytosolic reductases, and the role is assessed for these enzymes in reducing hydroxypyruvate and glyoxylate that may be leaked from peroxisomes.  相似文献   

6.
Myrosinase (β-thioglucoside glucohydrolase, EC 3.2.3.1) was purified to apparent homogeneity from light-grown cress (Lepidium sativum L.) seedlings. This enzyme, which catalyzes hydrolysis of the glucosinolate sinigrin (Km, 115 micromolar) at an optimum pH of 5.5 in sodium citrate buffer, had a native molecular weight of 130 ± 5 kilodaltons and an isoelectric point of 4.7 to 4.9. SDS-PAGE revealed two polypeptides with molecular weights of 62 and 65 kilodaltons. Both subunits contained carbohydrate as shown by periodic acid-Schiff staining. The purified enzyme hydrolyzed p-nitrophenyl-β-d-glucoside (Km, 2.0 millimolar) at an optimum pH of 6.5 in phosphate buffer. The indolizidine alkaloid castanospermine, a known inhibitor of O-glycosidases, competitively inhibited the hydrolyses of sinigrin (thioglucosidase activity) and p-nitrophenyl-β-d-glucoside (O-glucosidase activity) with Ki values of 5 and 6 micromolar, respectively. In contrast, the related polyhydroxyalkaloids swainsonine and deoxynojirimycin were without effect upon these hydrolyses.  相似文献   

7.
Glycolate oxidase (GO) has been identified in the endocyanom Cyanophora paradoxa which has peroxisome-like organelles and cyanelles instead of chloroplasts. The enzyme used or formed equimolar amounts of O2 or H2O2 and glyoxylate, respectively. Aerobically, the enzyme did not reduce the artificial electron acceptor dichlorophenol indophenol. However, after an inhibitor of glycolate dehydrogenase, KCN (2 millimolar), was added to the assay medium, considerable aerobic glycolate:dichlorophenol indophenol reductase activity was detectable. The leaf GO inhibitor 2-hydroxybutynoate (30 micromolar), which binds irreversibly to the flavin moiety of the active site of leaf GO, inhibited Cyanophora GO and pea (Pisum sativum L.) GO to the same extent. This suggests that the active sites of both enzymes are similar. Cyanophora GO and pea GO cannot oxidize d-lactate. In contrast to GO from pea or other organisms, the affinity of Cyanophora GO for l-lactate is very low (Km 25 millimolar). Another important difference is that Cyanophora GO produced sigmoidal kinetics with O2 as varied substrate, whereas pea GO produced normal Michaelis-Menten kinetics. It is concluded that there is considerable inhomogeneity among the glycolate-oxidizing enzymes from Cyanophora, pea, and other organisms. The specific catalase activity in Cyanophora was only one-tenth of that in leaves. NADH-and NADPH-dependent hydroxypyruvate reductase (HPR) and glyoxylate reductase activities were detected in Cyanophora. NADH-HPR was markedly inhibited by hydroxypyruvate above 0.5 millimolar. Variable substrate inhibition was observed with glyoxylate in homogenates from different algal cultures. It is proposed that Cyanophora has multiple forms of HPR and glyoxylate reductase, but no enzyme clearly resembling leaf peroxisomal HPR was identified in these homogenates. Moreover, no serine:glyoxylate aminotransferase activity was detected. These results collectively indicate the possibility that the glycolate metabolism in Cyanophora deviates from that in leaves.  相似文献   

8.
The activity of cytidine 5′-diphosphate (CDP) choline: 1,2-diacylglycerol cholinephosphotransferase (EC 2.7.8.2) in developing soybean (Glycine max L. var Williams 82) seeds was 3 to 5 times higher in cotyledons grown at 20°C than in those grown at 35°C. Some characteristics of the enzyme from cotyledons cultured at 20 and 35°C were compared. In preparations from both growth temperatures, the enzyme showed a pH optimum of 7, Km of 7.0 micromolar for CDP-choline, and an optimum assay temperature of 45°C. Both enzyme preparations were stimulated by increasing concentrations of Mg2+ or Mn2+, up to 10 millimolar and 50 micromolar, respectively, though Mn2+ produced lower activities than Mg2+. Enzymes from both 20 and 35°C show the same specificity for exogenous diacylglycerol. No metabolic effectors were detected by addition of heat treated extracts to the assay mixture. The above findings suggest that the higher enzyme activity at 20°C can be attributed to a higher level of the enzyme rather than to the involvement of isozymes or metabolic effectors. Enzyme activity decreased rapidly during culture at 35°C, indicating a rapid turnover of the enzyme. The level of temperature modulation was found to be a function of seed developmental stage.  相似文献   

9.
Glutamate synthase (GOGAT), a key enzyme in the pathway for the assimilation of symbiotically fixed dinitrogen (N2) into amino acids in alfalfa (Medicago sativa L.) root nodules, was purified and used to produce high titer polyclonal antibodies. Purification resulted in a 208-fold increase in specific activity to 13 micromole per minute per milligram of protein and an activity yield of 37%. Further purification to near homogeneity was achieved by fast protein liquid chromatography, but with substantial loss of activity. Enzymic activity was highly labile, losing 3% per hour even when substrates, stabilizers, and reducing agents were included in buffers. However, activity could be partially stabilized for up to 1 month by storing GOGAT at −80°C in 50% glycerol. The subunit molecular weight of GOGAT was estimated at 200 ± 7 kilodaltons with a native molecular weight of 235 ± 16 kilodaltons, which suggested that GOGAT is a monomer of unusually high molecular weight. The pl was estimated to be 6.6. The Km values for glutamine, α-ketoglutarate, and NADH were 466, 33, and 4.2 micromolar, respectively. Antibodies were produced to NADH-GOGAT. Specificity of the antibodies was shown by immunotitration of GOGAT activity. Alfalfa nodule NADH-GOGAT antibodies cross-reacted with polypeptides of a similar molecular weight in a number of legume species. Western blots probed with anti-GOGAT showed that the high GOGAT activity of nodules as compared to roots was associated with increased levels of GOGAT polypeptides. Nodule NADH-GOGAT appeared to be highly expressed in effective nodules and little if any in other organs.  相似文献   

10.
Microsomes from apical buds of pea (Pisum sativum L. var. Téléphone à rames) seedlings hydroxylate lauric acid at the ω-position. This oxidation is catalyzed by a cytochrome P-450 enzyme which differs from laurate hydroxylases previously described in microorganisms and mammals by its strict substrate specificity and the ability of low NADH concentrations to support unusually high oxidation rates. The apparent Km for lauric acid was 20 micromolar. NADPH- and NADH-dependent laurate hydroxylation followed non-Michaelian kinetics with apparent Km values ranging from 0.2 to 28 micromolar for NADPH, and 0.2 to 318 micromolar for NADH. When induced by the photomorphogenic photoreceptor phytochrome, the time course for the enhancement of laurate ω-hydroxylase was totally different from that of the cinnamic acid 4-hydroxylase, providing evidence for the existence of multiple cytochrome P-450 species in pea microsomes.  相似文献   

11.
The synthetic chelating agent EDTA can mobilize radionuclides and heavy metals in the environment. Biodegradation of EDTA should reduce this mobilization. Although several bacteria have been reported to mineralize EDTA, little is known about the biochemistry of EDTA degradation. Understanding the biochemistry will facilitate the removal of EDTA from the environment. EDTA-degrading activities were detected in cell extracts of bacterium BNC1 when flavin mononucleotide (FMN), NADH, and O2 were present. The degradative enzyme system was separated into two different enzymes, EDTA monooxygenase and an FMN reductase. EDTA monooxygenase oxidized EDTA to glyoxylate and ethylenediaminetriacetate (ED3A), with the coconsumption of FMNH2 and O2. The FMN reductase provided EDTA monooxygenase with FMNH2 by reducing FMN with NADH. The FMN reductase was successfully substituted in the assay mixture by other FMN reductases. EDTA monooxygenase was purified to greater than 95% homogeneity and had a single polypeptide with a molecular weight of 45,000. The enzyme oxidized both EDTA complexed with various metal ions and uncomplexed EDTA. The optimal conditions for activity were pH 7.8 and 35°C. Kms were 34.1 μM for uncomplexed EDTA and 8.5 μM for MgEDTA2−; this difference in Km indicates that the enzyme has greater affinity for MgEDTA2−. The enzyme also catalyzed the release of glyoxylate from nitrilotriacetate and diethylenetriaminepentaacetate. EDTA monooxygenase belongs to a small group of FMNH2-utilizing monooxygenases that attack carbon-nitrogen, carbon-sulfur, and carbon-carbon double bonds.  相似文献   

12.
Barley (Hordeum vulgare L.) endosperm from developing seeds was found to contain relatively high activities of cytosolic NAD(P)H-dependent hydroxypyruvate reductase (HPR-2) and isocitrate dehydrogenase (ICDH). In contrast, activities of peroxisomal NADH-dependent hydroxypyruvate reductase (HPR-1) and glycolate oxidase as well as cytosolic NAD(P)H-dependent glyoxylate reductase were very low or absent in the endosperm both during maturation and seed germination, indicating the lack of a complete glycolate cycle in this tissue. In addition, activities of cytosolic glucose-6-phosphate dehydrogenase and glyceraldehyde-3-phosphate dehydrogenase were low or absent in the endosperm. The endosperm HPR-2 exhibited similar properties to those of an earlier described HPR-2 from green leaves, e.g. activities with both hydroxypyruvate and glyoxylate, utilization of both NADPH and NADH as cofactors, and a strong uncompetitive inhibition by oxalate (Ki in the order of micromolar). In etiolated leaves, both HPR-1 and HPR-2 were present with the same activity as in green leaves, indicating that the lack of HPR-1 in the endosperm is not a general feature of non-photosynthetic tissues. We conclude that the endosperm has considerable capacity for cytosolic NADP/NADPH cycling via HPR-2 and ICDH, the former being possibly involved in the utilization of a serine-derived carbon.  相似文献   

13.
14.
Foley T  Beale SI 《Plant physiology》1982,70(5):1495-1502
γ,δ-Dioxovaleric acid (DOVA) has been proposed as a precursor to heme and chlorophyll in plants and algae. DOVA transaminase activity was found in extracts of the unicellular green alga Euglena gracilis Klebs strain Z Pringsheim. Optimum conversion of DOVA to δ-aminolevulinic acid (ALA) occurred at pH 6.8. ALA formation was linear with time for at least 30 minutes at 37° C and was proportional to amount of cell extract in the incubation mixture. Boiled cell extract was inactive. DOVA transaminase from either wild-type or aplastidic derivative strain W14ZNaIL ran as a single band in agarose gel permeation chromatography, with a calculated molecular weight of 98,000 ± 3,000. l-Glutamic acid was the most effective amino donor. d-Glutamic acid was inactive. Km values for l-glutamic acid and DOVA were 11 and 1.1 millimolar, respectively. Pyridoxal phosphate stimulated activity maximally at 30 micromolar, and (aminooxy)acetate was strongly inhibitory. Glyoxylic acid was a competitive inhibitor with respect to DOVA, with an inhibition constant of 0.62 millimolar. Wild-type and aplastidic cells vielded equal activity, 31 ± 1 nanomoles ALA per 30 minutes per 107 cells, whether grown in light or dark. DOVA transaminase could not be separated from glyoxylate transaminase activity by agarose gel permeation or diethylaminoethyl-cellulose column chromatography. In all fractions, glyoxylate transaminase activity was at least 75 times greater than DOVA transaminase activity. DOVA transamination appears to be catalyzed by glyoxylate transaminase, and not to be of physiological significance with respect to chlorophyll synthesis in Euglena.  相似文献   

15.
Bromphenol blue, which was reduced with dithionite, was found to support nitrate reduction catalyzed by squash NADH:nitrate reductase at a rate about 5 times greater than NADH with freshly prepared enzyme and 10 times or more with enzyme having been frozen and thawed. Kinetic analysis of bromphenol blue as a substrate for squash nitrate reductase yielded apparent Km values of 60 micromolar for bromphenol blue at 10 millimolar nitrate and 500 micromolar for nitrate at 0.2 millimolar bromphenol blue. With the same preparation of enzyme the apparent Km values were 9 micromolar for NADH at 10 millimolar nitrate and 50 micromolar nitrate at 0.1 millimolar NADH. Bromphenol blue was found to be a noncompetitive inhibitor versus NADH with a Ki of 0.3 millimolar. When squash NADH:nitrate reductase activity was inactivated with p-hydroxymercuribenzoate or denatured by heating at 40°C, the bromphenol blue nitrate reductase activity was not lost. These results were taken to indicate that bromphenol blue and NADH donated electrons to nitrate reductase at different sites. When monoclonal antibodies prepared against corn and squash nitrate reductases were used to inhibit the nitrate reductase activities supported by NADH, bromphenol blue, and methyl viologen, differential inhibition was found which tended to indicate that the three electron donors were interacting with the enzyme at different sites. One monoclonal antibody prepared against squash nitrate reductase inhibited all three activities of both corn and squash nitrate reductase. It appears this antibody may bind to a highly conserved antigenic site in the nitrate binding region of the enzyme.  相似文献   

16.
A K-12 strain of Escherichia coli that overproduces methylenetetrahydrofolate reductase (MetF) has been constructed, and the enzyme has been purified to apparent homogeneity. A plasmid specifying MetF with six histidine residues added to the C terminus has been used to purify histidine-tagged MetF to homogeneity in a single step by affinity chromatography on nickel-agarose, yielding a preparation with specific activity comparable to that of the unmodified enzyme. The native protein comprises four identical 33-kDa subunits, each of which contains a molecule of noncovalently bound flavin adenine dinucleotide (FAD). No additional cofactors or metals have been detected. The purified enzyme catalyzes the reduction of methylenetetrahydrofolate to methyltetrahydrofolate, using NADH as the reductant. Kinetic parameters have been determined at 15°C and pH 7.2 in a stopped-flow spectrophotometer; the Km for NADH is 13 μM, the Km for CH2-H4folate is 0.8 μM, and the turnover number under Vmax conditions estimated for the reaction is 1,800 mol of NADH oxidized min−1 (mol of enzyme-bound FAD)−1. NADPH also serves as a reductant, but exhibits a much higher Km. MetF also catalyzes the oxidation of methyltetrahydrofolate to methylenetetrahydrofolate in the presence of menadione, which serves as an electron acceptor. The properties of MetF from E. coli differ from those of the ferredoxin-dependent methylenetetrahydrofolate reductase isolated from the homoacetogen Clostridium formicoaceticum and more closely resemble those of the NADH-dependent enzyme from Peptostreptococcus productus and the NADPH-dependent enzymes from eukaryotes.  相似文献   

17.
The activities of three enzymes and the concentration of intermediates involved in the synthesis of N,N-dimethyltryptamine (DMT) from endogenous tryptophan (TRP) have been measured in vitro in seedlings of Phalaris aquatica L. cv Australian Commercial over 16 days after planting. The activities of tryptophan decarboxylase and the two N-methyl-transferases increased rapidly to maximal rates of substrate conversion at day 5 of 95, 1000, and 2200 micromoles per hour per milliliter, respectively. After these maximal rates, the activities decreased rapidly. The concentration of intermediates increased rapidly from zero in the seeds to maximal values of 25 and 53 micromolar at day 5 for tryptamine (T) and N-methyltryptamine (MT), respectively, 1000 micromolar at day 6 for TRP, and 650 micromolar at day 8 for DMT. The concentration of DMT and of all the intermediates in its synthesis declined rapidly after the maximal value had been reached. A mathematical model of the pathway from TRP to DMT using these enzymes correctly predicts the concentrations of T and MT, intermediates whose concentration is determined only by the pathway, and confirms that these three enzymes are responsible for the in vivo synthesis of DMT. Kinetic studies are reported for these enzymes. Tryptophan decarboxylase uses pyridoxal phosphate (PALP) as a coenzyme and has the following kinetic constants: KmPALP = 2.5 micromolar, KmTRP = 200 micromolar, KiMT = 5 millimolar, and KiDMT = 4 millimolar. The N-methyltransferases use S-adenosylmethionine (SAM) as substrate; S-adenosylhomocysteine (SAH) is assumed to be the product. The mechanism of secondary indolethylamine-N-methyltransferase, determined by initial velocity studies, is rapid equilibrium random with formation of both dead end complexes. Secondary indolethylamine-N-methyltransferase methylates both MT and 5-methoxy-N-methyltryptamine (5MeOMT). The kinetic constants for the methylation of MT are: KMT = 40 ± 6, KSAM = 55 ± 15, KDMT = 60, KSAH = 4.3 ± 0.4 micromolar with unity interaction factors. The kinetic constants for the conversion of 5MeOMT to 5-methoxy-N,N-dimethyltryptamine (5MeODMT) are K5MeOMT = 40 ± 10, KSAM = 90 ± 40, and KSAH = 2.9 ± 0.3 micromolar with unity interaction factors, except for SAM-5MeODMT = 2.0 ± 0.9 and SAH-5MeOMT = 0.45 ± 0.25. The kinetic constants for primary indolethylamine N-methyltransferase are KmT = 20, KmSAM = 40, KiDMT = 450 micromolar with the substrates binding independently.  相似文献   

18.
Glycine hydroxamate is a competitive inhibitor of glycine decarboxylation and serine formation (referred to as glycine decarboxylase activity) in particulate preparations obtained from both callus and leaf tissue of tobacco. In preparations from tobacco callus tissues, the Ki for glycine hydroxamate was 0.24 ± 0.03 millimolar and the Km for glycine was 5.0 ± 0.5 millimolar. The inhibitor was chemically stable during assays of glycine decarboxylase activity, but reacted strongly when incubated with glyoxylate. Glycine hydroxamate blocked the conversion of glycine to serine and CO2in vivo when callus tissue incorporated and metabolized [1-14C]glycine, [1-14C]glycolate, or [1-14C]glyoxylate. The hydroxamate had no effect on glyoxylate aminotransferase activities in vivo, and the nonenzymic reaction between glycine hydroxamate and glyoxylate did not affect the flow of carbon in the glycolate pathway in vivo. Glycine hydroxamate is the first known reversible inhibitor of the photorespiratory conversion of glycine to serine and CO2.  相似文献   

19.
Mitochondria isolated from pea (Pisum sativum L.) leaves are able to transport the keto acid, oxaloacetate, from the reaction medium into he mitochondrial matrix at high rates. The rate of uptake by the mitochondria was measured as the rate of disappearance of oxaloacetate from the reaction medium as it was reduced by matrix malate dehydrogenase using NADH provided by glycine oxidation. The oxaloacetate transporter was identifed as being distinct from the dicarboxylate and the α-ketoglutarate transporters because of its inhibitor sensitivities and its inability to interact with other potential substrates. Phthalonate and phthalate were competitive inhibitors of oxaloacetate transport with Ki values of 60 micromolar and 2 millimolar, respectively. Butylmalonate, an inhibitor of the dicarboxylate and α-ketoglutarate transporters, did not alter the rate of oxaloacetate transport. In addition, a 1000-fold excess of malate, malonate, succinate, α-ketoglutarate, or phosphate had little effect on the rate of oxaloacetate transport. The Km for the oxaloacetate transporter was about 15 micromolar with a maximum velocity of over 500 nanomoles per milligram mitochondrial protein/min at 25°C. No requirement for a counter ion to move against oxaloacetate was detected and the highest rates of uptake occurred at alkaline pH values. An equivalent transporter has not been reported in animal mitochondria.  相似文献   

20.
Zhao Z  Hu X  Ross CW 《Plant physiology》1987,84(4):987-988
To prepare tissues for analysis of NAD+, NADH, NADP+, and NADPH, common practice is to freeze samples in liquid nitrogen, often followed by freeze-drying, before extraction in HCl or NaOH. With cucumber (Cucumis sativus L.) cotyledons, prefreezing in liquid nitrogen or slower freezing to −20°C yielded substantially lower values for NADH and NADPH than obtained from samples homogenized immediately in acid or base. Freeze-drying after freezing in liquid nitrogen generally caused even lower values of those coenzymes. We suggest that direct extraction is more likely to yield accurate results with cotyledons and other plant parts.  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