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.
Anionic polysaccharides are of growing interest in the biotechnology industry due to their potential pharmaceutical applications in drug delivery and wound treatment. Chemical composition and polymer length strongly influence the physical and biological properties of the polysaccharide and thus its potential industrial and medical applications. One promising approach to determining monomer composition and controlling the degree of polymerization involves the use of polysaccharide lyases, which catalyze the depolymerization of anionic polysaccharides via a β-elimination mechanism. Utilization of these enzymes for the production of custom-made oligosaccharides requires a high degree of control over substrate specificity. Previously, we characterized a polysaccharide lyase (Smlt1473) from Stenotrophomonas maltophilia k279a, which exhibited significant activity against hyaluronan (HA), poly-β-d-glucuronic acid (poly-GlcUA), and poly-β-d-mannuronic acid (poly-ManA) in a pH-regulated manner. Here, we utilize a sequence structure guided approach based on a homology model of Smlt1473 to identify nine putative substrate-binding residues and examine their effect on substrate specificity via site-directed mutagenesis. Interestingly, single point mutations H221F and R312L resulted in increased activity and specificity toward poly-ManA and poly-GlcUA, respectively. Furthermore, a W171A mutant nearly eliminated HA activity, while increasing poly-ManA and poly-GlcUA activity by at least 35%. The effect of these mutations was analyzed by comparison with the high resolution structure of Sphingomonas sp. A1-III alginate lyase in complex with poly-ManA tetrasaccharide and by taking into account the structural differences between HA, poly-GlcUA, and poly-ManA. Overall, our results demonstrate that even minor changes in active site architecture have a significant effect on the substrate specificity of Smlt1473, whose structural plasticity could be applied to the design of highly active and specific polysaccharide lyases.  相似文献   

2.
Polysaccharide lyases (PLs) are a broad class of microbial enzymes that degrade anionic polysaccharides. Equally broad diversity in their polysaccharide substrates has attracted interest in biotechnological applications such as biomass conversion to value-added chemicals and microbial biofilm removal. Unlike other PLs, Smlt1473 present in the clinically relevant Stenotrophomonas maltophilia strain K279a demonstrates a wide range of pH-dependent substrate specificities toward multiple, diverse polysaccharides: hyaluronic acid (pH 5.0), poly-β-D-glucuronic (celluronic) acid (pH 7.0), poly-β-D-mannuronic acid, and poly-α-L-guluronate (pH 9.0). To decode the pH-driven multiple substrate specificities and selectivity in this single enzyme, we present the X-ray structures of Smlt1473 determined at multiple pH values in apo and mannuronate-bound states as well as the tetra-hyaluronate-docked structure. Our results indicate that structural flexibility in the binding site and N-terminal loop coupled with specific substrate stereochemistry facilitates distinct modes of entry for substrates having diverse charge densities and chemical structures. Our structural analyses of wild-type apo structures solved at different pH values (5.0–9.0) and pH-trapped (5.0 and 7.0) catalytically relevant wild-type mannuronate complexes (1) indicate that pH modulates the catalytic microenvironment for guiding structurally and chemically diverse polysaccharide substrates, (2) further establish that molecular-level fluctuation in the enzyme catalytic tunnel is preconfigured, and (3) suggest that pH modulates fluctuations resulting in optimal substrate binding and cleavage. Furthermore, our results provide key insight into how strategies to reengineer both flexible loop and regions distal to the active site could be developed to target new and diverse substrates in a wide range of applications.  相似文献   

3.
The poly-γ-d-glutamic acid capsule of Bacillus anthracis is a barrier to infection by B. anthracis-specific bacteriophages. Capsule expression was found to completely inhibit lytic infection by γ phage, an observation supported by the demonstration that this phage does not elaborate a hydrolase that would facilitate penetration through the protective capsule outer layer.  相似文献   

4.
Glucanase (endo-β-1, 3-glucan 3-glucanohydrolase, EC 3.2.1.6, laminarinase, callase) and chitinase (poly-β-1, 4-[2-acetamido-2-deoxy] -d-glucoside glycanohydrolase, EC 3.2.1.14) were extracted from ethylene-treated bean (Phaseolus vulgaris L. cv. Red Kidney) leaves and purified on hydroxyapatite and carboxymethyl Sephadex columns. The glucanase prepared was homogeneous as judged by analytical centrifugation data, electrophoresis, and antibody-antigen reactions. On the basis of gel filtration, antibody-antigen reactions, and amino acid analysis, the molecular weight was estimated to be between 11,500 and 12,500. However, ultracentrifugation gave a higher estimate of 34,000. The glucanase had an isoelectric point near pH 11 and was specific for β-1, 3-linkages. The chitinase was only partially purified as judged by electrophoretic behavior.  相似文献   

5.
Gum arabic (GA) is widely used as an emulsion stabilizer and coating in several industrial applications, such as foods and pharmaceuticals. GA contains a complex carbohydrate moiety, and the nonreducing ends of the side chains are often capped with l-rhamnose; thus, enzymes that can remove these caps are promising tools for the structural analysis of the carbohydrates comprising GA. In this study, GA-specific l-rhamnose-α-1,4-d-glucuronate lyase from the fungus Fusarium oxysporum 12S (FoRham1) was cloned and characterized. FoRham1 showed the highest amino acid sequence similarity with enzymes belonging to the glycoside hydrolase family 145; however, the catalytic residue on the posterior pocket of the β-propeller fold protein was not conserved. The catalytic residues of FoRham1 were instead conserved with ulvan lyases belonging to polysaccharide lyase family 24. Kinetic analysis showed that FoRham1 has the highest catalytic efficiency for the substrate α-l-rhamnose-(1→4)-d-glucuronic acid. The crystal structures of ligand-free and α-l-rhamnose-(1→4)-d-glucuronic acid –bound FoRham1 were determined, and the active site was identified on the anterior side of the β-propeller. The three-dimensional structure of the active site and mutagenesis analysis revealed the detailed catalytic mechanism of FoRham1. Our findings offer a new enzymatic tool for the further analysis of the GA carbohydrate structure and for elucidating its physiological functions in plants. Based on these results, we renamed glycoside hydrolase family 145 as a new polysaccharide lyase family 42, in which FoRham1 is included.  相似文献   

6.
Marine microbes use alginate lyases to degrade and catabolize alginate, a major cell wall matrix polysaccharide of brown seaweeds. Microbes frequently contain multiple, apparently redundant alginate lyases, raising the question of whether these enzymes have complementary functions. We report here on the molecular cloning and functional characterization of three exo-type oligoalginate lyases (OalA, OalB, and OalC) from Vibrio splendidus 12B01 (12B01), a marine bacterioplankton species. OalA was most active at 16°C, had a pH optimum of 6.5, and displayed activities toward poly-β-d-mannuronate [poly(M)] and poly-α-l-guluronate [poly(G)], indicating that it is a bifunctional enzyme. OalB and OalC were most active at 30 and 35°C, had pH optima of 7.0 and 7.5, and degraded poly(M·G) and poly(M), respectively. Detailed kinetic analyses of oligoalginate lyases with poly(G), poly(M), and poly(M·G) and sodium alginate as substrates demonstrated that OalA and OalC preferred poly(M), whereas OalB preferred poly(M·G). The catalytic efficiency (kcat/Km) of OalA against poly(M) increased with decreasing size of the substrate. OalA showed kcat/Km from 2,130 mg−1 ml s−1 for the trisaccharide to 224 mg−1 ml s−1 for larger oligomers of ∼50 residues, and 50.5 mg−1 ml s−1 for high-molecular-weight alginate. Although OalA was most active on the trisaccharide, OalB and OalC preferred dimers. Taken together, our results indicate that these three Oals have complementary substrate scopes and temperature and pH adaptations.  相似文献   

7.
The metabolism of myo-inositol-2-14C, d-glucuronate-1-14C, d-glucuronate-6-14C, and l-methionine-methyl-14C to cell wall polysaccharides was investigated in excised root-tips of 3 day old Zea mays seedlings. From myo-inositol, about one-half of incorporated label was recovered in ethanol insoluble residues. Of this label, about 90% was solubilized by treatment, first with a preparation of pectinase-EDTA, then with dilute hydrochloric acid. The only labeled constituents in these hydrolyzates were d-galacturonic acid, d-glucuronic acid, 4-O-methyl-d-glucuronic acid, d-xylose, and l-arabinose, or larger oligosaccharide fragments containing these units. Medium external to excised root-tips grown under sterile conditions in myo-inositol-2-14C contained labeled polysaccharide.  相似文献   

8.
Pseudomonas aeruginosa is an opportunistic pathogen that forms chronic biofilm infections in the lungs of cystic fibrosis patients. A major component of the biofilm during these infections is the exopolysaccharide alginate, which is synthesized at the inner membrane as a homopolymer of 1–4-linked β-d-mannuronate. As the polymer passages through the periplasm, 22–44% of the mannuronate residues are converted to α-l-guluronate by the C5-epimerase AlgG to produce a polymer of alternating β-d-mannuronate and α-l-guluronate blocks and stretches of polymannuronate. To understand the molecular basis of alginate epimerization, the structure of Pseudomonas syringae AlgG has been determined at 2.1-Å resolution, and the protein was functionally characterized. The structure reveals that AlgG is a long right-handed parallel β-helix with an elaborate lid structure. Functional analysis of AlgG mutants suggests that His319 acts as the catalytic base and that Arg345 neutralizes the acidic group during the epimerase reaction. Water is the likely catalytic acid. Electrostatic surface potential and residue conservation analyses in conjunction with activity and substrate docking studies suggest that a conserved electropositive groove facilitates polymannuronate binding and contains at least nine substrate binding subsites. These subsites likely align the polymer in the correct register for catalysis to occur. The presence of multiple subsites, the electropositive groove, and the non-random distribution of guluronate in the alginate polymer suggest that AlgG is a processive enzyme. Moreover, comparison of AlgG and the extracellular alginate epimerase AlgE4 of Azotobacter vinelandii provides a structural rationale for the differences in their Ca2+ dependence.  相似文献   

9.
Alginate lyases are enzymes that degrade alginate through β-elimination of the glycosidic bond into smaller oligomers. We investigated the alginate lyases from Vibrio splendidus 12B01, a marine bacterioplankton species that can grow on alginate as its sole carbon source. We identified, purified, and characterized four polysaccharide lyase family 7 alginates lyases, AlyA, AlyB, AlyD, and AlyE, from V. splendidus 12B01. The four lyases were found to have optimal activity between pH 7.5 and 8.5 and at 20 to 25°C, consistent with their use in a marine environment. AlyA, AlyB, AlyD, and AlyE were found to exhibit a turnover number (kcat) for alginate of 0.60 ± 0.02 s−1, 3.7 ± 0.3 s−1, 4.5 ± 0.5 s−1, and 7.1 ± 0.2 s−1, respectively. The Km values of AlyA, AlyB, AlyD, and AlyE toward alginate were 36 ± 7 μM, 22 ± 5 μM, 60 ± 2 μM, and 123 ± 6 μM, respectively. AlyA and AlyB were found principally to cleave the β-1,4 bonds between β-d-mannuronate and α-l-guluronate and subunits; AlyD and AlyE were found to principally cleave the α-1,4 bonds involving α-l-guluronate subunits. The four alginate lyases degrade alginate into longer chains of oligomers.  相似文献   

10.
An α-l-arabinofuranosidase has been purified 1043-fold from radish (Raphanus sativus L.) seeds. The purified enzyme was a homogeneous glycoprotein consisting of a single polypeptide with an apparent molecular weight of 64,000 and an isoelectric point value of 4.7, as evidenced by denaturing gel electrophoresis and reversed-phase or size-exclusion high-performance liquid chromatography and isoelectric focusing. The enzyme characteristically catalyzes the hydrolysis of p-nitrophenyl α-l-arabinofuranoside and p-nitrophenyl β-d-xylopyranoside in a constant ratio (3:1) of the initial velocities at pH 4.5, whereas the corresponding α-l-arabinopyranoside and β-d-xylofuranoside are unsusceptible. The following evidence was provided to support that a single enzyme with one catalytic site was responsible for the specificity: (a) high purity of the enzyme preparation, (b) an invariable ratio of the activities toward the two substrates throughout the purification steps, (c) a parallelism of the activities in activation with bovine serum albumin and in heat inactivation of the enzyme as well as in the inhibition with heavy metal ions and sugars such as Hg2+, Ag+, l-arabino-(1→4)-lactone, and d-xylose, and (d) results of the mixed substrate kinetic analysis using the two substrates. The enzyme was shown to split off α-l-arabinofuranosyl residues in sugar beet arabinan, soybean arabinan-4-galactan, and radish seed and leaf arabinogalactan proteins. Arabinose and xylose were released by the action of the enzyme on oat-spelt xylan. Synergistic action of α-l-arabinofuranosidase and β-d-galactosidase on radish seed arabinogalactan protein resulted in the extensive degradation of the carbohydrate moiety.  相似文献   

11.
Cells of yellow-poplar (Liriodendron tulipifera L.) were transformed by direct gene transfer and regenerated into plants by somatic embryogenesis. Plasmid DNA bearing marker genes encoding β-glucuronidase (GUS) and neomycin phosphotransferase (NPT II) were introduced by microprojectile bombardment into single cells and small cell clusters isolated from embryogenic suspension cultures. The number of full-length copies of the GUS gene in independently transformed callus lines ranged from approximately 3 to 30. An enzyme-linked immunosorbent assay for NPT II and a fluorometric assay for GUS showed that the expression of both enzymes varied by less than fourfold among callus lines. A histochemical assay for GUS activity revealed a heterogeneous pattern of staining with the substrate 5-bromo-4-chloro-3-indoyl-β-d-glucuronic acid in some transformed cell cultures. However, cell clusters reacting positively (blue) or negatively (white) with 5-bromo-4-chloro-3-indoyl-β-d-glucuronic acid demonstrated both GUS activity and NPT II expression in quantitative assays. Somatic embryos induced from transformed cell cultures were found to be uniformly GUS positive by histochemical analysis. All transgenic plants sampled expressed the two marker genes in both root and shoot tissues. GUS activity was found to be higher in leaves than roots by fluorometric and histochemical assays. Conversely, roots expressed higher levels of NPT II than leaves.  相似文献   

12.
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   

13.
Fry SC  Northcote DH 《Plant physiology》1983,73(4):1055-1061
Cultured spinach (Spinacia oleracea L. cv Monstrous Viroflay) cells incorporated exogenous l-[3H]arabinose sequentially into β-l-arabinopyranose-1-phosphate, uridine diphospho-β-l-arabinopyranose, uridine diphospho-α-d-xylopyranose and (in some experiments) α-d-xylopyranose-1-phosphate. The amount of 3H in each of these compounds reached a plateau after a few minutes, and could be rapidly chased with nonradioactive l-arabinose, demonstrating rapid turnover. After a few minutes' lag, incorporation of 3H into the arabinofuranosyl, arabinopyranosyl, and xylopyranosyl residues of polysaccharides was linear with respect to time. The kinetics of labeling were compatible with UDP-β-l-arabinopyranose and UDP-α-d-xylopyranose being the immediate precursors of arabians (both the pyranose and the furanose residues) and xylans, respectively. No other radioactive nucleotides were formed; in particular, UDP-arabinofuranose was absent. There was no evidence for conversion of arabinopyranose to arabinofuranose within the polysaccharides, suggesting that this conversion occurs during polymer synthesis. The glycolipids detected showed too slow a turnover to be intermediates of pentosan synthesis.  相似文献   

14.
Lyases cleave glycosaminoglycans (GAGs) in an eliminative mechanism and are important tools for the structural analysis and oligosaccharide preparation of GAGs. Various GAG lyases have been identified from terrestrial but not marine organisms even though marine animals are rich in GAGs with unique structures and functions. Herein we isolated a novel GAG lyase for the first time from the marine bacterium Vibrio sp. FC509 and then recombinantly expressed and characterized it. It showed strong lyase activity toward hyaluronan (HA) and chondroitin sulfate (CS) and was designated as HA and CS lyase (HCLase). It exhibited the highest activities to both substrates at pH 8.0 and 0.5 m NaCl at 30 °C. Its activity toward HA was less sensitive to pH than its CS lyase activity. As with most other marine enzymes, HCLase is a halophilic enzyme and very stable at temperatures from 0 to 40 °C for up to 24 h, but its activity is independent of divalent metal ions. The specific activity of HCLase against HA and CS reached a markedly high level of hundreds of thousands units/mg of protein under optimum conditions. The HCLase-resistant tetrasaccharide Δ4,5HexUAα1-3GalNAc(6-O-sulfate)β1-4GlcUA(2-O-sulfate)β1-3GalNAc(6-O-sulfate) was isolated from CS-D, the structure of which indicated that HCLase could not cleave the galactosaminidic linkage bound to 2-O-sulfated d-glucuronic acid (GlcUA) in CS chains. Site-directed mutagenesis indicated that HCLase may work via a catalytic mechanism in which Tyr-His acts as the Brønsted base and acid. Thus, the identification of HCLase provides a useful tool for HA- and CS-related research and applications.  相似文献   

15.
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   

16.
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   

17.
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   

18.
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   

19.
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   

20.
Hoson T  Nevins DJ 《Plant physiology》1989,90(4):1353-1358
Antiserum was raised against the Avena sativa L. caryopsis β-d-glucan fraction with an average molecular weight of 1.5 × 104. Polyclonal antibodies recovered from the serum after Protein A-Sepharose column chromatography precipitated when cross-reacted with high molecular weight (1→3), (1→4)-β-d-glucans. These antibodies were effective in suppression of cell wall autohydrolytic reactions and auxin-induced decreases in noncellulosic glucose content of the cell wall of maize (Zea mays L.) coleoptiles. The results indicate antibody-mediated interference with in situ β-d-glucan degradation. The antibodies at a concentration of 200 micrograms per milliliter also suppress auxin-induced elongation by about 40% and cell wall loosening (measured by the minimum stress-relaxation time of the segments) of Zea coleoptiles. The suppression of elongation by antibodies was imposed without a lag period. Auxin-induced elongation, cell wall loosening, and chemical changes in the cell walls were near the levels of control tissues when segments were subjected to antibody preparation precipitated by a pretreatment with Avena caryopsis β-d-glucans. These results support the idea that the degradation of (1→3), (1→4)-β-d-glucans by cell wall enzymes is associated with the cell wall loosening responsible for auxin-induced elongation.  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