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 46 毫秒
1.
Cell suspension cultures of parsley (Petroselinum crispum) accumulated coumarin phytoalexins and exhibited increased β-1,3-glucanase activity when treated with either a purified α-1,4-d-endopolygalacturonic acid lyase from Erwinia carotovora or oligogalacturonides solubilized from parsley cell walls by endopolygalacturonic acid lyase. Coumarin accumulation induced by the plant cell wall elicitor was preceded by increases in the activities of phenylalanine ammonia lyase (PAL), 4-coumarate:CoA ligase (4CL) and S-adenosyl-l-methionine:xanthotoxol O-methyltransferase (XMT). The time courses for the changes in these three enzyme activities were similar to those observed in cell cultures treated with a fungal glucan elicitor. The plant cell wall elicitor was found to act synergistically with the fungal glucan elicitor in the induction of coumarin phytoalexins. As much as a 10-fold stimulation in coumarin accumulation above the calculated additive response was observed in cell cultures treated with combinations of plant and fungal elicitors. The synergistic effect was also observed for the induction of PAL, 4CL, and XMT activities. These results demonstrate that plant cell wall elicitors induce at least two distinct biochemical responses in parsley cells and further support the role of oligogalacturonides as important regulators of plant defense.  相似文献   

2.
Elicitors of phytoalexin accumulation in soybean (Glycine max L. Merr., cv Wayne) cotyledons were released from soybean cell walls and from citrus pectin by partial acid hydrolysis. These two hydrolysates yielded nearly identical distributions of elicitor activity when fractionated on anion-exchange columns. Chromatography of the pectin elicitor on gel filtration and high-pressure anion-exchange columns did not further purify the elicitor. Elicitor activity of the preparation was lost by treatment with either endo-α-1,4-polygalacturonase or pectate lyase. Glycosyl residue compositions of the purified elicitors from cell walls and pectin were both found to be approximately 98% galacturonosyl residues. Linkage analysis of the pectin elicitor showed that most, if not all, of the galacturonosyl residues were α-1,4-linked. The high-mass molecular ions detected by fast atom bombardment-mass spectrometry of the most active elicitor fractions from cell walls and pectin both corresponded precisely to a molecule composed of 12 galacturonosyl residues. These results suggest that dodeca-α-1,4-d-galacturonide is the active elicitor, but the possibility remains that the active component could be a slightly modified oligogalacturonide present, but not detected, in the purified fractions.  相似文献   

3.
Root hairs provide a model system to study plant cell growth, yet little is known about the polysaccharide compositions of their walls or the role of these polysaccharides in wall expansion. We report that Arabidopsis thaliana root hair walls contain a previously unidentified xyloglucan that is composed of both neutral and galacturonic acid–containing subunits, the latter containing the β-d-galactosyluronic acid-(1→2)-α-d-xylosyl-(1→ and/or α-l-fucosyl-(1→2)-β-d-galactosyluronic acid-(1→2)-α-d-xylosyl-(1→) side chains. Arabidopsis mutants lacking root hairs have no acidic xyloglucan. A loss-of-function mutation in At1g63450, a root hair–specific gene encoding a family GT47 glycosyltransferase, results in the synthesis of xyloglucan that lacks galacturonic acid. The root hairs of this mutant are shorter than those of the wild type. This mutant phenotype and the absence of galacturonic acid in the root xyloglucan are complemented by At1g63450. The leaf and stem cell walls of wild-type Arabidopsis contain no acidic xyloglucan. However, overexpression of At1g63450 led to the synthesis of galacturonic acid–containing xyloglucan in these tissues. We propose that At1g63450 encodes XYLOGLUCAN-SPECIFIC GALACTURONOSYLTRANSFERASE1, which catalyzes the formation of the galactosyluronic acid-(1→2)-α-d-xylopyranosyl linkage and that the acidic xyloglucan is present only in root hair cell walls. The role of the acidic xyloglucan in root hair tip growth is discussed.  相似文献   

4.
Ruminococcus albus is a typical ruminal bacterium digesting cellulose and hemicellulose. Cellobiose 2-epimerase (CE; EC 5.1.3.11), which converts cellobiose to 4-O-β-d-glucosyl-d-mannose, is a particularly unique enzyme in R. albus, but its physiological function is unclear. Recently, a new metabolic pathway of mannan involving CE was postulated for another CE-producing bacterium, Bacteroides fragilis. In this pathway, β-1,4-mannobiose is epimerized to 4-O-β-d-mannosyl-d-glucose (Man-Glc) by CE, and Man-Glc is phosphorolyzed to α-d-mannosyl 1-phosphate (Man1P) and d-glucose by Man-Glc phosphorylase (MP; EC 2.4.1.281). Ruminococcus albus NE1 showed intracellular MP activity, and two MP isozymes, RaMP1 and RaMP2, were obtained from the cell-free extract. These enzymes were highly specific for the mannosyl residue at the non-reducing end of the substrate and catalyzed the phosphorolysis and synthesis of Man-Glc through a sequential Bi Bi mechanism. In a synthetic reaction, RaMP1 showed high activity only toward d-glucose and 6-deoxy-d-glucose in the presence of Man1P, whereas RaMP2 showed acceptor specificity significantly different from RaMP1. RaMP2 acted on d-glucose derivatives at the C2- and C3-positions, including deoxy- and deoxyfluoro-analogues and epimers, but not on those substituted at the C6-position. Furthermore, RaMP2 had high synthetic activity toward the following oligosaccharides: β-linked glucobioses, maltose, N,N′-diacetylchitobiose, and β-1,4-mannooligosaccharides. Particularly, β-1,4-mannooligosaccharides served as significantly better acceptor substrates for RaMP2 than d-glucose. In the phosphorolytic reactions, RaMP2 had weak activity toward β-1,4-mannobiose but efficiently degraded β-1,4-mannooligosaccharides longer than β-1,4-mannobiose. Consequently, RaMP2 is thought to catalyze the phosphorolysis of β-1,4-mannooligosaccharides longer than β-1,4-mannobiose to produce Man1P and β-1,4-mannobiose.  相似文献   

5.
Previous results (TJ Buckhout, Planta [1989] 178: 393-399) indicated that the structural specificity of the H+-sucrose symporter on the plasma membrane from sugar beet leaves (Beta vulgaris L.) was specific for the sucrose molecule. To better understand the structural features of the sucrose molecule involved in its recognition by the symport carrier, the inhibitory activity of a variety of phenylhexopyranosides on sucrose uptake was tested. Three competitive inhibitors of sucrose uptake were found, phenyl-α-d-glucopyranoside, phenyl-α-d-thioglucopyranoside, and phenyl-α-d-4-deoxythioglucopyranoside (PDTGP; Ki = 67, 180, and 327 micromolar, respectively). The Km for sucrose uptake was approximately 500 micromolar. Like sucrose, phenyl-α-d-thioglucopyranoside and to a lesser extent, PDTGP induced alkalization of the external medium, which indicated that these derivatives bound to and were transported by the sucrose symporter. Phenyl-α-d-3-deoxy-3-fluorothioglucopyranoside, phenyl-α-d-4-deoxy-4-fluorothioglucopyranoside, and phenyl-α-d-thioallopyranoside only weakly but competively inhibited sucrose uptake with Ki values ranging from 600 to 800 micromolar, and phenyl-α-d-thiomannopyranoside, phenyl-β-d-glucopyranoside, and phenylethyl-β-d-thiogalactopyranoside did not inhibit sucrose uptake. Thus, the hydroxyl groups of the fructose portion of sucrose were not involved in a specific interaction with the carrier protein because phenyl and thiophenyl derivatives of glucose inhibited sucrose uptake and, in the case of phenyl-α-d-thioglucopyranoside and PDTGP, were transported.  相似文献   

6.
Chloroplasts prepared from mesophyll protoplasts of the primary leaf of wheat (Triticum aestivum L. cv Egret) contain about 50% of the cellular β-galactosidase (EC 3.2.1.23) activity. More than 80% of this activity is associated with the stroma and most of the remainder, although tightly bound to the thylakoids, can be washed free with sodium pyrophosphate. The vacuole contained about 20% and the remaining enzyme was presumed to be cytoplasmic or associated with one of the other organelles. Both the vacuolar and chloroplast enzymes were capable of releasing galactose from the galactolipid monogalactosyldiacylglycerol. Apart from their distinct locations within the cells, we conclude that the enzymes are different because they differed with respect to assay pH-optimum, comparative activity against the synthetic substrates phenyl-β-d-galactoside, 4-methylumbelliferyl-β-d-galactoside, 6-bromo-2-naphthyl-β-d-galactoside, the disaccharide lactose, and the inhibitors d-galactose and d-galactono-1,4-lactone.  相似文献   

7.
Enzymes of sucrose breakdown in soybean nodules: alkaline invertase   总被引:4,自引:4,他引:0  
Morell M  Copeland L 《Plant physiology》1984,74(4):1030-1034
The specific activities of acid and alkaline invertases (β-d-fructofuranoside fructohydrolase, EC 3.2.1.26), sucrose synthase (UDPglucose: d-fructose 2-α-d-glucosyltransferase, EC 2.4.1.13), hexokinase (ATP: d-hexose 6-phosphotransferase, EC 2.7.1.1), and fructokinase (ATP: d-fructose 6-phosphotransferase, EC 2.7.1.4) were determined in soybean (Glycine max L. Merr cv Williams) nodules at different stages of development and, for comparison, in roots of nonnodulated soybeans. Alkaline invertase and sucrose synthase were both involved in sucrose metabolism in the nodules, but there was only a small amount of acid invertase present. The nodules contained more phosphorylating activity with fructose than glucose. Essentially all of the alkaline invertase, sucrose synthase, and fructokinase were in the soluble fraction of nodule extracts whereas hexokinase was in the bacteroid, plant particulate, and soluble fractions.  相似文献   

8.
1. The previous study (Conchie, Gelman & Levvy, 1967b) of the specificity of β-glucosidase, β-galactosidase and β-d-fucosidase in barley, limpet, almond emulsin and rat epididymis was extended to α-l-arabinosidase. 2. The inhibitory action of l-arabinono-(1→5)-lactone was tested against all four types of enzyme, and α-l-arabinosidase was examined for inhibition by glucono-, galactono- and d-fucono-lactone. 3. In emulsin, the enzyme that hydrolyses β-glucosides, β-galactosides and β-d-fucosides also hydrolyses α-l-arabinosides. Rat epididymis resembles emulsin except that, as already noted, it lacks β-glucosidase activity. 4. In the limpet, α-l-arabinosidase activity is associated with the enzyme that hydrolyses β-glucosides and β-d-fucosides, and not with the separate β-galactosidase. 5. The effects of the different lactones on the barley preparation suggest that α-l-arabinosidase activity is associated with the β-galactosidase rather than with the enzyme that hydrolyses β-glucosides and β-d-fucosides. Fractionation and heat-inactivation experiments indicate that there is also a separate α-l-arabinosidase in the preparation.  相似文献   

9.
1. In barley, β-glucosidase and β-galactosidase are separate enzymes. The former also displays β-d-fucosidase activity. 2. In the limpet, Patella vulgata, β-glucosidase activity is associated with the β-d-fucosidase, previously shown to be a separate entity from the β-galactosidase also present. 3. Almond emulsin presents all three activities as a single enzyme. Each is equally inhibited by glucono-, galactono- and d-fucono-lactone. 4. In rat epididymis, there is no significant β-glucosidase activity, nor is there appreciable inhibition of the β-galactosidase and β-d-fucosidase activities of the preparation by gluconolactone.  相似文献   

10.
A gene cluster involved in N-glycan metabolism was identified in the genome of Bacteroides thetaiotaomicron VPI-5482. This gene cluster encodes a major facilitator superfamily transporter, a starch utilization system-like transporter consisting of a TonB-dependent oligosaccharide transporter and an outer membrane lipoprotein, four glycoside hydrolases (α-mannosidase, β-N-acetylhexosaminidase, exo-α-sialidase, and endo-β-N-acetylglucosaminidase), and a phosphorylase (BT1033) with unknown function. It was demonstrated that BT1033 catalyzed the reversible phosphorolysis of β-1,4-d-mannosyl-N-acetyl-d-glucosamine in a typical sequential Bi Bi mechanism. These results indicate that BT1033 plays a crucial role as a key enzyme in the N-glycan catabolism where β-1,4-d-mannosyl-N-acetyl-d-glucosamine is liberated from N-glycans by sequential glycoside hydrolase-catalyzed reactions, transported into the cell, and intracellularly converted into α-d-mannose 1-phosphate and N-acetyl-d-glucosamine. In addition, intestinal anaerobic bacteria such as Bacteroides fragilis, Bacteroides helcogenes, Bacteroides salanitronis, Bacteroides vulgatus, Prevotella denticola, Prevotella dentalis, Prevotella melaninogenica, Parabacteroides distasonis, and Alistipes finegoldii were also suggested to possess the similar metabolic pathway for N-glycans. A notable feature of the new metabolic pathway for N-glycans is the more efficient use of ATP-stored energy, in comparison with the conventional pathway where β-mannosidase and ATP-dependent hexokinase participate, because it is possible to directly phosphorylate the d-mannose residue of β-1,4-d-mannosyl-N-acetyl-d-glucosamine to enter glycolysis. This is the first report of a metabolic pathway for N-glycans that includes a phosphorylase. We propose 4-O-β-d-mannopyranosyl-N-acetyl-d-glucosamine:phosphate α-d-mannosyltransferase as the systematic name and β-1,4-d-mannosyl-N-acetyl-d-glucosamine phosphorylase as the short name for BT1033.  相似文献   

11.
β-Galactosidase-catalysed hydrolysis of β-d-galactopyranosyl azide   总被引:3,自引:3,他引:0  
1. β-d-Galactopyranosyl azide is hydrolysed by the β-galactosidase of Escherichia coli to galactose and azide ion at a mechanistically significant rate. 2. Methyl 1-thio-β-d-galactopyranoside is a competitive inhibitor of the hydrolysis of the azide and of o-nitrophenyl β-d-galactopyranoside with Ki 1.8mm. 3. β-Galactosidase can thus hydrolyse a range of substrates of general structure β-d-galactopyranosyl-X(Y), where the atom X has a lone pair of electrons on which the enzyme may act as a Lewis or Brønsted acid, but in which the length of the bond cleaved varies significantly, which is inconsistent with the orbital steering hypothesis.  相似文献   

12.
Manganese superoxide dismutase (MnSOD), a foremost antioxidant enzyme, plays a key role in angiogenesis. Barley-derived (1.3) β-d-glucan (β-d-glucan) is a natural water-soluble polysaccharide with antioxidant properties. To explore the effects of β-d-glucan on MnSOD-related angiogenesis under oxidative stress, we tested epigenetic mechanisms underlying modulation of MnSOD level in human umbilical vein endothelial cells (HUVECs) and angiogenesis in vitro and in vivo. Long-term treatment of HUVECs with 3% w/v β-d-glucan significantly increased the level of MnSOD by 200% ± 2% compared to control and by 50% ± 4% compared to untreated H2O2-stressed cells. β-d-glucan-treated HUVECs displayed greater angiogenic ability. In vivo, 24 hrs-treatment with 3% w/v β-d-glucan rescued vasculogenesis in Tg (kdrl: EGFP) s843Tg zebrafish embryos exposed to oxidative microenvironment. HUVECs overexpressing MnSOD demonstrated an increased activity of endothelial nitric oxide synthase (eNOS), reduced load of superoxide anion (O2) and an increased survival under oxidative stress. In addition, β-d-glucan prevented the rise of hypoxia inducible factor (HIF)1-α under oxidative stress. The level of histone H4 acetylation was significantly increased by β-d-glucan. Increasing histone acetylation by sodium butyrate, an inhibitor of class I histone deacetylases (HDACs I), did not activate MnSOD-related angiogenesis and did not impair β-d-glucan effects. In conclusion, 3% w/v β-d-glucan activates endothelial expression of MnSOD independent of histone acetylation level, thereby leading to adequate removal of O2, cell survival and angiogenic response to oxidative stress. The identification of dietary β-d-glucan as activator of MnSOD-related angiogenesis might lead to the development of nutritional approaches for the prevention of ischemic remodelling and heart failure.  相似文献   

13.
Bacteria which were β-d-galactosidase and β-d-glucuronidase positive or expressed only one of these enzymes were isolated from environmental water samples. The enzymatic activity of these bacteria was measured in 25-min assays by using the fluorogenic substrates 4-methylumbelliferyl-β-d-galactoside and 4-methylumbelliferyl-β-d-glucuronide. The enzyme activity, enzyme induction, and enzyme temperature characteristics of target and nontarget bacteria in assays aimed at detecting coliform bacteria and Escherichia coli were investigated. The potential interference of false-positive bacteria was evaluated. Several of the β-d-galactosidase-positive nontarget bacteria but none of the β-d-glucuronidase-positive nontarget bacteria contained unstable enzyme at 44.5°C. The activity of target bacteria was highly inducible. Nontarget bacteria were induced much less or were not induced by the inducers used. The results revealed large variations in the enzyme levels of different β-d-galactosidase- and β-d-glucuronidase-positive bacteria. The induced and noninduced β-d-glucuronidase activities of Bacillus spp. and Aerococcus viridans were approximately the same as the activities of induced E. coli. Except for some isolates identified as Aeromonas spp., all of the induced and noninduced β-d-galactosidase-positive, noncoliform isolates exhibited at least 2 log units less mean β-d-galactosidase activity than induced E. coli. The noncoliform bacteria must be present in correspondingly higher concentrations than those of target bacteria to interfere in the rapid assay for detection of coliform bacteria.Indicators of pollution (e.g., coliforms, fecal coliforms, and Escherichia coli) are traditionally used for monitoring the microbiological safety of water supplies and recreational water. Several techniques for detection of coliforms and E. coli are based on enzymatic hydrolysis of fluorogenic or chromogenic substrates for β-d-galactosidase and β-d-glucuronidase (9, 20). Current methods of recovery are usually culture based, and the analysis time is 18 to 24 h. In addition to enzymatic activity, these techniques use growth at appropriate temperatures in the presence of inhibitors, combined with demonstration of enzymatic activity, to selectively detect target bacteria.Rapid methods which require less than 6 h and are based on chromogenic, fluorogenic, or chemiluminogenic substrates for detection of coliforms, fecal coliforms, or E. coli have been described (13, 10, 27, 28). These rapid assays are based on the assumption that β-d-galactosidase and β-d-glucuronidase are markers for coliforms and E. coli, respectively. However, when the incubation time is 1 h or less, growth is not a selective step, and all β-d-galactosidase-positive or β-d-glucuronidase-positive microorganisms in a water sample contribute to the activity measured. At low initial concentrations of target bacteria (i.e., E. coli and total coliforms), increasing the preincubation time to 5 to 6 h did not result in a predominance of target bacteria compared to nontarget bacteria (28).The β-d-galactosidase or β-d-glucuronidase activity calculated per cultivable coliform or fecal coliform bacterium in environmental samples can be 1 to 2 log units higher than the activity per induced E. coli cell in pure culture (11, 26). The presence of active, noncultivable bacteria can be one reason for this. Studies of survival (7, 24, 25) and disinfection (26) of E. coli have shown that loss of cultivability does not necessarily result in a loss of β-d-galactosidase activity. The presence of false-positive bacteria can be another reason.β-d-Galactosidase has been found in numerous microorganisms, including gram-negative bacteria (e.g., strains belonging to the Enterobacteriaceae, Vibrionaceae, Pseudomonadaceae, and Neisseriaceae), several gram-positive bacteria, yeasts, protozoa, and fungi (17, 29). β-d-Glucuronidase is produced by most E. coli strains and also by other members of the Enterobacteriaceae, including some Shigella and Salmonella strains and a few Yersinia, Citrobacter, Edwardia, and Hafnia strains. Production of β-d-glucuronidase by Flavobacterium spp., Bacteroides spp., Staphylococcus spp., Streptococcus spp., anaerobic corynebacteria, and Clostridium has also been reported (12).High numbers of false-positive bacteria in sewage and contaminated water have been revealed by enumeration of β-d-galactosidase- and β-d-glucuronidase-positive CFU on nonselective agar supplemented with fluorogenic or chromogenic substrates (11, 28). Whether the activity from nontarget organisms can be neglected in a rapid assay depends on the number of nontarget organisms compared with the number of target bacteria and also on the level of their enzyme activity. Plant and algal biomass must be present at high concentrations to interfere in rapid bacterial β-d-galactosidase and β-d-glucuronidase assays (8).The main objective of this study was to investigate the enzyme characteristics of β-d-galactosidase- and β-d-glucuronidase-positive bacteria isolated from environmental water samples and to evaluate the potential influence of false-positive bacteria in rapid assays for coliform bacteria or E. coli in water. The effect of temperature on enzyme activity and on the interference of nontarget bacteria in the rapid assays was investigated as an important factor.(Some of the results were presented at the 97th General Meeting of the American Society for Microbiology 1997, Miami Beach, Fla., 4 to 8 May 1997.)  相似文献   

14.
Accumulation of d-leucine, d-allo-isoleucine, and d-valine was observed in the growth medium of a lactic acid bacterium, Lactobacillus otakiensis JCM 15040, and the racemase responsible was purified from the cells and identified. The N-terminal amino acid sequence of the purified enzyme was GKLDKASKLI, which is consistent with that of a putative γ-aminobutyrate aminotransferase from Lactobacillus buchneri. The putative γ-aminobutyrate aminotransferase gene from L. buchneri JCM 1115 was expressed in recombinant Escherichia coli and then purified to homogeneity. The enzyme catalyzed the racemization of a broad spectrum of nonpolar amino acids. In particular, it catalyzed at high rates the epimerization of l-isoleucine to d-allo-isoleucine and d-allo-isoleucine to l-isoleucine. In contrast, the enzyme showed no γ-aminobutyrate aminotransferase activity. The relative molecular masses of the subunit and native enzyme were estimated to be about 49 kDa and 200 kDa, respectively, indicating that the enzyme was composed of four subunits of equal molecular masses. The Km and Vmax values of the enzyme for l-isoleucine were 5.00 mM and 153 μmol·min−1·mg−1, respectively, and those for d-allo-isoleucine were 13.2 mM and 286 μmol·min−1·mg−1, respectively. Hydroxylamine and other inhibitors of pyridoxal 5′-phosphate-dependent enzymes completely blocked the enzyme activity, indicating the enzyme requires pyridoxal 5′-phosphate as a coenzyme. This is the first evidence of an amino acid racemase that specifically catalyzes racemization of nonpolar amino acids at the C-2 position.  相似文献   

15.
The mode of inhibition of UDP, one of the products of the reaction catalyzed by (1→3)-β-d-glucan synthase in sugar beet (Beta vulgaris L.) was investigated. In the absence of added UDP, the enzyme, in the presence of Ca2+, Mg2+, and cellobiose, exhibited Michaelis-Menten kinetics and had an apparent Km of 260 micromolar for UDP-glucose. Complex effects on the kinetics of the (1→3)-β-d-glucan synthase were observed in the presence of UDP. At high UDP-glucose concentrations, i.e. greater than the apparent Km, UDP behaved as a competitive inhibitor with an apparent Ki of 80 micromolar. However, at low UDP-glucose concentrations, reciprocal plots of enzyme activity versus substrate concentration deviated sharply from linearity. This unusual effect of UDP is similar to that reported for fungal (1→3)-β-d-glucan synthase. However, papulacandin B, a potent inhibitor of this fungal enzyme, had no effect on the plant (1→3)-β-d-glucan synthase isolated from sugar beet petioles. The inhibitory effect of UDP was also compared with other known inhibitors of glucan synthases.  相似文献   

16.
1. A number of disaccharides and oligosaccharides have been isolated from the products of mild acid hydrolysis of the specific substance from Lactobacillus casei, serological group C. 2. The major disaccharide is O-β-d-glucopyranosyl-(1→3)-N-acetyl- d-galactosamine (B4) and evidence is presented for the structure of a tetrasaccharide composed of O-β-d-glucopyranosyl-(1→6)-d-galactose (B1) joined through its reducing end group to B4. 3. Disaccharide B1 is also a component of a trisaccharide O-β-d-glucopyranosyl-(1→6)-O-β- d-galactopyranosyl-(1→6)-N-acetyl-d-glucosamine (A7). 4. A number of other oligosaccharides have been shown to be related structurally. 5. The ability of certain of the oligosaccharides to inhibit the precipitin reaction has been studied. The disaccharide B1 is more effective as an inhibitor than gentiobiose and the trisaccharide A7 is considerably more effective than B1. 6. These results have been compared with those obtained previously for the composition of the cell wall.  相似文献   

17.
Dwarf maize (Zea mays L.), a mutant deficient in gibberellin synthesis, provides an excellent model to study the influence of gibberellin on biochemical processes related to plant development. Alterations in the chemical structure of the cell wall mediated by gibberellin were examined in seedlings of this mutant. The composition of the walls of roots, mesocotyl, coleoptile, and primary leaves of dwarf maize was similar to that of normal maize and other cereal grasses. Glucuronoarabinoxylans constituted the principal hemicelluloses, but walls also contained substantial amounts of xyloglucan and mixed-linkage β-d-glucan. Root growth in dwarf maize was essentially normal, but growth of mesocotyl and primary leaves was severely retarded. Injection of the gibberellin into the cavity of the coleoptile resulted in a marked increase in elongation of the primary leaves. This elongation was accompanied by increases in total wall mass, but the proportion of β-d-glucan decreased from 20% to 15% of the hemicellulosic polysaccharide. During leaf expansion, the proportion decreased further to only 10%. Through 4 days incubation, the proportion of β-d-glucan in leaves of control seedlings without gibberellin was nearly constant. Extraction of exo- and endo-β-d-glucan hydrolases from purified cell walls and assay against a purified oat bran β-d-glucan demonstrated that gibberellin increased the activity of the endo-β-d-glucan hydrolase. These and other data support the hypothesis that β-d-glucan metabolism is central to control of cell expansion in cereal grasses.  相似文献   

18.
The core lipopolysaccharide (LPS) of Aeromonas hydrophila AH-3 and Aeromonas salmonicida A450 is characterized by the presence of the pentasaccharide α-d-GlcN-(1→7)-l-α-d-Hep-(1→2)-l-α-d-Hep-(1→3)-l-α-d-Hep-(1→5)-α-Kdo. Previously it has been suggested that the WahA protein is involved in the incorporation of GlcN residue to outer core LPS. The WahA protein contains two domains: a glycosyltransferase and a carbohydrate esterase. In this work we demonstrate that the independent expression of the WahA glycosyltransferase domain catalyzes the incorporation of GlcNAc from UDP-GlcNAc to the outer core LPS. Independent expression of the carbohydrate esterase domain leads to the deacetylation of the GlcNAc residue to GlcN. Thus, the WahA is the first described bifunctional glycosyltransferase enzyme involved in the biosynthesis of core LPS. By contrast in Enterobacteriaceae containing GlcN in their outer core LPS the two reactions are performed by two different enzymes.  相似文献   

19.
β-Primeverosidase (PD) is a disaccharide-specific β-glycosidase in tea leaves. This enzyme is involved in aroma formation during the manufacturing process of oolong tea and black tea. PD hydrolyzes β-primeveroside (6-O-β-d-xylopyranosyl-β-d-glucopyranoside) at the β-glycosidic bond of primeverose to aglycone, and releases aromatic alcoholic volatiles of aglycones. PD only accepts primeverose as the glycone substrate, but broadly accepts various aglycones, including 2-phenylethanol, benzyl alcohol, linalool, and geraniol. We determined the crystal structure of PD complexes using highly specific disaccharide amidine inhibitors, N-β-primeverosylamidines, and revealed the architecture of the active site responsible for substrate specificity. We identified three subsites in the active site: subsite −2 specific for 6-O-β-d-xylopyranosyl, subsite −1 well conserved among β-glucosidases and specific for β-d-glucopyranosyl, and wide subsite +1 for hydrophobic aglycone. Glu-470, Ser-473, and Gln-477 act as the specific hydrogen bond donors for 6-O-β-d-xylopyranosyl in subsite −2. On the other hand, subsite +1 was a large hydrophobic cavity that accommodates various aromatic aglycones. Compared with aglycone-specific β-glucosidases of the glycoside hydrolase family 1, PD lacks the Trp crucial for aglycone recognition, and the resultant large cavity accepts aglycone and 6-O-β-d-xylopyranosyl together. PD recognizes the β-primeverosides in subsites −1 and −2 by hydrogen bonds, whereas the large subsite +1 loosely accommodates various aglycones. The glycone-specific activity of PD for broad aglycone substrates results in selective and multiple release of temporally stored alcoholic volatile aglycones of β-primeveroside.  相似文献   

20.
The exact subcellular location of sucrose synthase (UDP-d-glucose: d-fructose 2-α-d-glucosyltransferase, EC 2.4.1.13) in Helianthus tuberosus tubers was studied by comparison of its activity in protoplasts with that of vacuoles isolated from them. Assuming 100% of the β-N-acetylglucosaminidase activity to be of vacuolar origin, less than 5% of both the sucrose synthase activity and the extravacuolar marker NAD-malate dehydrogenase was detected in the vacuole preparations. Sucrose synthase is therefore an extravacuolar enzyme. Its role in the inulin metabolism of H. tuberosus is discussed.  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