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.
Polyamine analogues have been studied as potential inhibitors or substrates of barley leaf polyamine oxidase. NH2(CH2)3NH(CH2)10NH2 was particularly effective as an inhibitor of spermine oxidation at pH 4·5 (Ki = 5 × 10?6 M). Methylglyoxal-bis(guanylhydrazone) inhibited spermine oxidation only slightly (Ki = 10?4 M). Activity with the polyamine analogues as substrates was generally 10% or less of the activity with spermine. The Km for oxygen was 3 × 10?4 M. The Km for spermine oxidation was independent of oxygen concentration. Using the N-methyl-2-benzothiazolone hydrazine reagent, 1-(3-aminopropyl)pyrroline was shown to be formed stoichiometrically by the enzyme on oxidation of spermine. The enzyme will not function as a dehydrogenase in the presence of oxygen with either potassium ferricyanide or dichlorophenolindophenol as electron acceptors. Activity in the leaves increased with age, up to 4 weeks. In the leaves of 11-week-old plants activity was lower than in leaves of 1-week-old plants. The enzyme was mainly associated with an easily-sedimented particulate fraction, and relatively small proportions were found in the cell wall or soluble fractions.  相似文献   

2.
Beside a cardinal role in coordination of many developmental processes in the plant, the phytohormone auxin has been recognized as a regulator of plant defense. The molecular mechanisms involved are still largely unknown. Using a sensitive chemiluminescence assay, which measures the oxidation of luminol in the presence of H2O2 by horseradish peroxidase (HRP), we report here on the ability of exogenously added indole-3-acetic acid (IAA) to enhance the suppressive effect of the root endophyte Piriformospora indica on the chitin-elicited oxidative burst in barley roots. Thus, the potential of P. indica to produce free IAA during the early colonization phase in barley might provide the symbiont with a means to interfere with the microbe-associated molecular patterns (MAMP)-triggered immunity.  相似文献   

3.
The polyamine oxidase of barley shoots is associated with a particle which sediments in low centrifugal fields. The enzyme was removed from these particles by washing in 0·5 M NaCl and then purified about 24-fold. The purified enzyme oxidized spermine stoicheiometrically to 1,3-diaminopropane and 1-(3-aminopropyl)pyrroline (pH optimum 4·0). Spermidine was oxidized to 1,3-diaminopropane and 1-pyrroline (pH optimum 6·6). At their respective pH optima, spermine is oxidized about 30 times faster than spermidine. Hydrogen peroxide was formed in the course of the polyamine oxidation. The enzyme was not sensitive to several copper chelating reagents but 2-hydroxyethylhydrazine caused 50% inhibition at 5 × 10−4 M. The enzyme was also present in particles in the roots of barley seedlings and in extracts of the leaves of oats, maize, rye and wheat.  相似文献   

4.
The oxidation of linoleic acid in incubation mixtures containing extracts of barley lipoxygenase and hydroperoxide isomerase, and the production of these enzymes in quiescent and germinated barley, were investigated. The ratio of 9-hydroperoxylinoleic acid to 13-hydroperoxylinoleic acid was higher for incubation mixtures containing extracts of quiescent barley than for mixtures containing extracts of germinated barley; production of 13-hydroperoxylinoleic acid from germinated barley exceeded that of quiescent barley. Hydroperoxy metabolites of linoleic acid were converted to 9-hydroxy-10-oxo-cis-12-octadecenoic acid, 13-hydroxy-10-oxo-trans-11-octadecenoic acid, and small amounts of 11-hydroxy-12,13-epoxy-cis-9-octadecenoic acid and 11-hydroxy-9,10-epoxy-cis-13-octadecenoic acid whether quiescent or germinated barley was the enzyme source; a fifth product, 13-hydroxy-12-oxo-cis-9-octadecenoic acid was formed only when germinated barley was the enzyme source.  相似文献   

5.
The pH optimum for the stability of the barley leaf polyamine oxidase is 4.8, which is also the pH optimum for its activity with spermine as substrate. Zonal centrifugation indicates that the enzyme is associated with a particle which is slightly more dense than chloroplasts, and the peak of activity corresponds with the peak of nucleic acid. Neither DNase nor RNase released the enzyme from the particles, despite the hydrolysis of more than 50% of the nucleic acid. The enzyme from the leaves of oat seedlings grown in the dark was purified 900-fold. Mg2+ and Ca2+ inhibited both barley and oat enzymes by ca 50% at 50 mM. The optimum pH for both spermine and spermidine oxidation by the oat enzyme was 6.5. The MW of the enzyme from both sources determined by gel chromatography was ca 85 000.  相似文献   

6.
Bowen JE 《Plant physiology》1976,57(3):353-357
At 2 C, all boron accumulated by excised barley roots (Hordeum vulgare L. cv. Herta) remains in the free space; i.e. active uptake is nil at this temperature. Three component fractions of free space B were apparent: (a) a surface contaminant film of B on blotted roots, (b) water free space B, and (c) B reversibly bound in the cell walls. A stoichiometric release of H+ from the roots in the presence of B indicated that B was bound by borate complexes with polysaccharides in the cell walls. Polysaccharide-borate complexes are much less stable than those of monosaccharides, and the bound B fraction could be readily removed by rinsing the roots in the presence of a monomeric polyol possessing the necessary cis-diol configuration. Cell wall material separated from excised barley roots had a B binding capacity 66% greater than that of intact roots.  相似文献   

7.
L-Tyrosine carboxy-lyase (E.C. 4. 1. 1. 25) was isolated fromroots of germinating barley (Hordeum vulgare). The enzyme requirespyridoxal phosphate for maximum activity. The optimum pH foractivity is about 7.0. The enzyme is inhibited by p-chloromercuribenzoateand hydroxylamine at 10–3 M. Enzyme activity is foundin extracts from young roots, especially from those in earlystages of development, but not in extracts from shoots of thesame plant. Localization and changes in the amounts of L-tyrosinecarboxy-lyase and aromatic amines in developing barley seedlingswere measured. Participation of carboxy-lyase in the formationof aromatic amines in barley roots is suggested. (Received July 17, 1970; )  相似文献   

8.
Addition of catecholamines at micromolar concentrations caused a dramatic stimulation of growth of tobacco (Nicotiana tabacum) thin cell layers (TCLs) and Acmella oppositifolia “hairy” root cultures. A threefold increase in the rate of ethylene evolution was observed in the catecholamine-treated explants. Aminooxyacetic acid and silver thiosulfate, inhibitors of ethylene biosynthesis and action, respectively, reduced the growth-promoting effect of dopamine. However, these compounds alone could also inhibit the growth of the TCL explants. When ethylene in the culture vessel was depleted by trapping with mercuric perchlorate, dopamine-stimulated growth was still obtained, suggesting that ethylene does not mediate the dopamine effect. Dopamine potentiated the growth of TCLs grown in Murashige and Skoog medium supplemented with indoleacetic acid (IAA) and kinetin. When IAA was replaced by 2,4-dichlorophenoxyacetic acid, dopamine addition showed no growth-promoting effect. Instead, 2,4-dichlorophenoxyacetic acid stimulated the growth of TCL explants to the same extent as that obtained with IAA plus dopamine. Because synthetic auxins do not appear to be substrates for IAA oxidizing enzymes, we hypothesized that catecholamines exert their effect by preventing IAA oxidation. Consistent with this explanation, dopamine (25 micromolar) inhibited IAA oxidase activity by 60 to 100% in crude enzyme extracts from tobacco roots and etiolated corn coleoptiles, but had no effect on peroxidase activity in the same extracts. Furthermore, addition of dopamine to TCL cultures resulted in a fourfold reduction in the oxidative degradation of [1-14C]IAA fed to the explants. Because the growth enhancement by catecholamines is observed in both IAA-requiring and IAA-independent cultures, we suggest that these aromatic amines may have a role in the regulation of IAA levels in vivo.  相似文献   

9.
Using Affigel Blue and oxamate-agarose affinity chromatography, lactate dehydrogenase (LDH) was purified 2000-fold from hypoxically induced barley roots. Molecular weights of the native and sodium dodecyl sulfate-denatured LDH protein were 157 and 40 kilodaltons, respectively, indicating a tetramer. Purified barley LDH was very similar in size and kinetic properties to potato LDH. However, their amino acid compositions differed substantially and antibodies raised against barley LDH did not cross-react with potato LDH on immunoblots, implying that the barley and potato LDHs are not closely related proteins. In vivo [35S] methionine labeling and immunoprecipitation experiments indicated that hypoxia increased the rate of LDH protein synthesis, and immunoblot analysis showed that LDH protein levels rose during hypoxia. We conclude that increased enzyme synthesis plays a major part in the induction of LDH enzyme activity by low O2 levels in barley roots.  相似文献   

10.
An enzyme present in extracts of the shoots of barley seedlings has been shown to synthesize coumarylagmatine from p-coumaryl-coenzyme A and [U-14C]agmatine.  相似文献   

11.
Igamberdiev AU  Bykova NV  Hill RD 《Planta》2006,223(5):1033-1040
NADH-dependent NO scavenging in barley extracts is linked to hemoglobin (Hb) expression and is inhibited by SH-reagents. Barley Hb has a single cysteine residue. To determine whether this cysteine was critical for NO scavenging, barley Hb and a mutated version, in which the single Cys79 was replaced by Ser, were over-expressed in Escherichia coli and purified to near homogeneity. The purified proteins exhibited very low NO-scavenging activity (12–14 nmol min−1 mg−1 protein) in the presence of NADH or NADPH. This activity was insensitive to SH-reagents. Addition of an extract from barley roots to either of the purified proteins resulted in high NADH-dependent NO turnover in a reaction that was sensitive to SH-reagents. A protein was purified from barley roots and identified by mass-spectrometry analysis as a cytosolic monodehydroascorbate reductase. It efficiently supported NADH-dependent NO scavenging in the presence of either native or mutated barley Hb. Ascorbate strongly facilitated the rate of metHb reduction. The K m for Hb was 0.3 μM, for ascorbate 0.6 mM and for NADH 4 μM. The reaction in the presence of monodehydroascorbate reductase was sensitive to SH-reagents with either form of the Hb. We conclude that metHb reduction and NO turnover do not involve direct participation of the Cys79 residue of barley Hb. NO scavenging is facilitated by monodehydroascorbate reductase mediating a coupled reaction involving ferric Hb reduction in the presence of ascorbate and NADH.  相似文献   

12.
IAA Oxidase Inhibitors from Normal and Mutant Maize Plants   总被引:1,自引:1,他引:0       下载免费PDF全文
Extracts of maize (Zea mays L.) plants contain substances which, in vitro, inhibit an indoleacetic acid (IAA) oxidase enzyme from maize. The extracts can be freed of inhibitors by dialysis or by passage through columns of polyvinylpyrrolidone powder. Inhibitor-free extracts contain an IAA oxidase enzyme which requires a phenolic co-factor and is stimulated by Mn2+.  相似文献   

13.
Rapid induction of na/h exchange activity in barley root tonoplast   总被引:12,自引:2,他引:10       下载免费PDF全文
Na+/H+ exchange activity in barley (Hordeum vulgare cv CM-72) root tonoplast was induced by Na+ even in the presence of inhibitors of protein synthesis. Induction occurred with a half-time of only 15 minutes. When salt-treated roots were transferred to a nutrient solution containing no Na+, the activity disappeared with a similar time course. The data suggest that Na+/H+ exchange was due to activation of an existing protein rather than to de novo protein synthesis.  相似文献   

14.
The gene HvNHX3 encoding a new isoform of vacuolar Na+/H+-antiporter was identified in barley. This gene is expressed in roots and leaves of barley seedlings, and it encodes a protein consisting of 541 amino acid residues with pre-dicted molecular weight 59.7 kDa. It was found that by its amino acid sequence HvNHX3 is closest to the Na+/H+-antiporter HbNHX1 of wild type from Hordeum brevisibulatum that grows on salt-marsh (solonchak) soils (95% homology). The expression of HvNHX3 during salt stress is increased several-fold in roots and leaves of barley seedlings. At the same time, the amount of HvNHX3 protein in roots does not change, but in leaves it increases significantly. It was shown using HvNHX3 immunolocalization in roots that this protein is present in all tissues, but in control plants it was clustered and in experimental plants after salt stress it was visualized as small granules. It has been proposed that HvNHX3 is converted into active form during declusterization. The conversion of HvNHX3 into its active form along with its quantitative increase in leaves during salt stress activates Na+/H+-exchange across the vacuolar membrane and Na+ release from cytoplasm, and, as a consequence, an increase of salt stress tolerance.  相似文献   

15.
Glycine betaine (GB) is a compatible solute accumulated by many plants under various abiotic stresses. GB is synthesized in two steps, choline → betaine aldehyde → GB, where a functional choline-oxidizing enzyme has only been reported in Amaranthaceae (a chloroplastic ferredoxin-dependent choline monooxygenase) thus far. Here, we have cloned a cDNA encoding a choline monooxygenase (CMO) from barley (Hordeum vulgare) plants, HvCMO. In barley plants under non-stress condition, GB had accumulated in all the determined organs (leaves, internodes, awn and floret proper), mostly in the leaves. The expression of HvCMO protein was abundant in the leaves, whereas the expression of betaine aldehyde dehydrogenase (BADH) protein was abundant in the awn, floret proper and the youngest internode than in the leaves. The accumulation of HvCMO mRNA was increased by high osmotic and low-temperature environments. Also, the expression of HvCMO protein was increased by the presence of high NaCl. Immunofluorescent labeling of HvCMO protein and subcellular fractionation analysis showed that HvCMO protein was localized to peroxisomes. [14C]choline was oxidized to betaine aldehyde and GB in spinach (Spinacia oleracea) chloroplasts but not in barley, which indicates that the subcellular localization of choline-oxidizing enzyme is different between two plant species. We investigated the choline-oxidizing reaction using recombinant HvCMO protein expressed in yeast (Saccharomyces cerevisiae). The crude extract of HvCMO-expressing yeast coupled with recombinant BBD2 protein converted [14C]choline to GB when NADPH was added as a cofactor. These results suggest that choline oxidation in GB synthesis is mediated by a peroxisomal NADPH-dependent choline monooxygenase in barley plants.  相似文献   

16.
α-Hydroxypyridine methane sulphonic acid (HPMS), isonicotinyl hydrazide (INH) and nialamide inhibit chlorophyll synthesis in etiolated barley leaves exposed to light. HPMS lowered the rate of protochlorophyllide regeneration but had little effect on the synthesis of protochlorophyll (P630) from exogenous δ-aminolaevulinic acid (ALA). The addition of glycine to HPMS treated leaves partially overcame the inhibition of chlorophyll synthesis. Glycine-[14C] was readily incorporated into ALA in dark-grown leaves. HPMS treatment increased the sp. act. of ALA in leaves fed glycine-[14C]. Glycollate oxidation was lower in extracts from HPMS treated leaves. Plants may therefore have two pathways for ALA production with the glutamate pathway becoming more important in conditions where photorespiration is high.  相似文献   

17.
In barley genotype susceptible toErysiphe graminis f. sp.hordei 1,3-β-glucan synthase activity in whole leaf extracts was higher in comparison with healthy plants. A positive correlation was found between the activity of 1,3-β-glucan synthase and the degree of barley resistance. On the contrary, the 1,3-β-D-glucanase activity in whole leaves was negatively correlated to host plant resistance. This phenomen is evident only in the early phase of plant pathogen interaction. However, in epidermal cells the 1,3-glucanase activity was not significantly changed after attack and the 1,3-glucan synthase activity was practically zero. Chitinase activity in inoculated leaves and epidermis was higher than in healthy ones, but no unambigous correlation was found between the enzyme activity and host resistance.  相似文献   

18.
Lactate dehydrogenase (LDH) activity in attached roots of barley and other cereals increased up to 20-fold during several days of severe hypoxia, reaching a maximum of about 2 micromoles per minute per gram fresh weight. In barley, induction of LDH activity was significant at 2.6% O2 and greatest at 0.06%, the lowest O2 concentration tested. Upon return to aerobic conditions, induced LDH activity declined with an apparent half-life of 2 days. The isozyme profile of barley LDH comprised 5 bands, consistent with a tetrameric enzyme with subunits encoded by two different Ldh genes. Changes in staining intensity of the isozymes as a function of O2 level suggested that one Ldh gene was preferentially expressed in severe hypoxia. When tracer [U-14C]glucose was supplied to induced roots under hypoxic conditions, lactate acquired label, but much less than either ethanol or alanine. Most of the [14C] lactate was secreted into the medium, whereas most other labeled anionic products were retained in the root. Neither hypoxic induction of LDH, nor lactate secretion by induced roots, is predicted from the Davies-Roberts hypothesis, which holds that lactate glycolysis ceases soon after the onset of hypoxia due to acidosis brought about by lactate accumulation in the cytoplasm. These results imply a functional significance for LDH beyond that assigned it in this hypothesis.  相似文献   

19.
In the present study, we have demonstrated that membrane-free extracts of etiolated shoots of Phaseolus coccineus seedlings show tocopherol oxidase activity. For this reaction, presence of membrane lipids, such as lecithin and mixture of plant lipids was required. The rate of the reaction was the highest for α-tocopherol and decreased in the order α ? β > γ > δ tocopherols. In the case of α-tocopherol, the main oxidation product was α-tocopherolquinone, while for the other tocopherol homologues the dominant products were other derivatives. When the enzyme activity was measured in leaves, hypocotyls and roots of etiolated seedlings of P. coccineus, the oxidase activity was the highest in extracts of leaves and decreased towards the roots where no activity was detected. The effect of hydrogen peroxide and of different inhibitors on the reaction suggest that tocopherol oxidase does not belong to peroxidases or flavin oxidases but rather to multi-copper oxidases, such as polyphenol oxidases or laccases. On the other hand, catechol, the well-known substrate of polyphenol oxidases and laccases, was not oxidized by the enzyme, indicating a high substrate specificity of the tocopherol oxidase.  相似文献   

20.
The present communication elucidates initially the topographic distribution of selenium in barley grains. Then by the fluorimetric method the uptake of selenium (selenite) in 8–16 d old germinating barley was estimated. Finally by means of75Se the anabolic and catabolic rates (turnover) of75Se (selenite) was compared. The distribution of selenium in barley was evaluated after microdissection of barley grains. In dried grains the highest concentration was found in husk and pericarp with about 0.6 ppm Se. Then followed Scutellum with 0.4 and 0.3 ppm in embryon. The aleurone layer, embryonic leaves, and initial root did only have 0.2 ppm Se. In order to know more about the uptake and distribution of selenium in 8-d-old barley, the plants were cultivated for a further 8 d in the culture medium with variation in selenite concentration. In roots and leaves, the uptake did not arrive at saturation during the period studied since the dose-response curve increased up to 0.34 mM selenite in the medium, whereas the selenium levels were about 200 ppm in roots and 30 ppm in leaves. However, the uptake was linear, with concentration during 8 d of cultivation up to 0.84 μM selenite for grain and stem. At higher concentrations the dose-response curve diminished its slope. At 0.34 mM selenite the concentration in grain increased to 6.87 ppm and in the stem to 8.13 ppm. The uptake, distribution, and catabolic rate of selenium components in germinating barley were further evaluated by exposing the plants to 0.0492 μCi75Se (12.6 μM selenite) for up to 4 d. Then the plants were moved to a selenium deficient medium for further 4 d. Then finally the medium was supplemented with high doses of cold selenite (0.126 mM selenite) for further 4 d. The first third period made it possible to estimate the rate of uptake. It was highest in roots (313 fmol/h/mg dw), i.e., about 10 times those of grains, stems, and leaves. The intermediate period where the barley was transferred to a selenium deficient medium made it possible to estimate the kinetics and eventual sparing mechanisms. The selenium losses were highest for leaves (39%), then followed by roots and stems (22 and 25%, respectively). The losses were lowest in grain with 9% Se losses. The losses were three times more pronounced during the first day than in the following 3 d. These data may argue that the selenium is distributed into different pools and that sparing mechanisms may be in function. The last period, i.e., the chase experiment, revealed the rate of elimination of selenium under conditions with surplus selenium. The catabolic rate was about 10 times faster in roots (169 fmol/h/mg dw) than in grains and about 8 times faster than in leaves.  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