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.
d-Galactan I is a polysaccharide with the disaccharide repeat unit structure [→3-β-d-Galf-(1→3)-α-d-Galp-(1→]. This glycan represents the lipopolysaccharide O antigen found in many Gram-negative bacteria, including several Klebsiella pneumoniae O serotypes. The polysaccharide is synthesized in the cytoplasm prior to its export via an ATP-binding cassette transporter. Sequence analysis predicts three galactosyltransferases in the d-galactan I genetic locus. They are WbbO (belonging to glycosyltransferase (GT) family 4), WbbM (GT-family 8), and WbbN (GT-family 2). The WbbO and WbbM proteins are each predicted to contain two domains, with the GT modules located toward their C termini. The N-terminal domains of WbbO and WbbM exhibit no similarity to proteins with known function. In vivo complementation assays suggest that all three glycosyltransferases are required for d-galactan I biosynthesis. Using a bacterial two-hybrid system and confirmatory co-purification strategies, evidence is provided for protein-protein interactions among the glycosyltransferases, creating a membrane-located enzyme complex dedicated to d-galactan I biosynthesis.  相似文献   

2.
Streptococcus pneumoniae is a Gram-positive human pathogen with a complex lipoteichoic acid (pnLTA) structure. Because the current structural model for pnLTA shows substantial inconsistencies, we reinvestigated purified and, more importantly, O-deacylated pnLTA, which is most suitable for NMR spectroscopy and electrospray ionization-MS spectrometry. We analyzed pnLTA of nonencapsulated pneumococcal strains D39Δcps and TIGR4Δcps, respectively. The data obtained allowed us to (re)define (i) the position and linkage of the repeating unit, (ii) the putative α-GalpNAc substitution at the ribitiol 5-phosphate (Rib-ol-5-P), and (iii) the length of (i.e. the number of repeating units in) the pnLTA chain. We here also describe for the first time that the terminal sugar residues in the pnLTA (Forssman disaccharide; α-d-GalpNAc-(1→3)-β-d-GalpNAc-(1→)), responsible for the cross-reactivity with anti-Forssman antigen antibodies, can be heterogeneous with respect to its degree of phosphorylcholine substitution in both O-6-positions. To assess the proinflammatory potency of pnLTA, we generated a (lipopeptide-free) Δlgt mutant of strain D39Δcps, isolated its pnLTA, and showed that it is capable of inducing IL-6 release in human mononuclear cells, independent of TLR2 activation. This finding was quite in contrast to LTA of the Staphylococcus aureus SA113Δlgt mutant, which did not activate human mononuclear cells in our experiments. Remarkably, this is also contrary to various other reports showing a proinflammatory potency of S. aureus LTA. Taken together, our study refines the structure of pnLTA and indicates that pneumococcal and S. aureus LTAs differ not only in their structure but also in their bioactivity.  相似文献   

3.
The Group B Streptococcus capsular polysaccharide type IX was isolated and purified, and the structure of its repeating unit was determined. Type IX capsule →4)[NeupNAc-α-(2→3)-Galp-β-(1→4)-GlcpNAc-β-(1→6)]-β-GlcpNAc-(1→4)-β-Galp-(1→4)-β-Glcp-(1→ appears most similar to types VII and V, although it contains two GlcpNAc residues. Genetic analysis identified differences in cpsM, cpsO, and cpsI gene sequences as responsible for the differentiation between the three capsular polysaccharide types, leading us to hypothesize that type V emerged from a recombination event in a type IX background.  相似文献   

4.
Bud dormancy is an adaptive strategy that perennials use to survive unfavorable conditions. Gentians (Gentiana), popular alpine flowers and ornamentals, produce overwintering buds (OWBs) that can persist through the winter, but the mechanisms regulating dormancy are currently unclear. In this study, we conducted targeted metabolome analysis to obtain clues about the metabolic mechanisms involved in regulating OWB dormancy. Multivariate analysis of metabolite profiles revealed metabolite patterns characteristic of dormant states. The concentrations of gentiobiose [β-d-Glcp-(1→6)-d-Glc] and gentianose [β-d-Glcp-(1→6)-d-Glc-(1→2)-d-Fru] significantly varied depending on the stage of OWB dormancy, and the gentiobiose concentration increased prior to budbreak. Both activation of invertase and inactivation of β-glucosidase resulted in gentiobiose accumulation in ecodormant OWBs, suggesting that gentiobiose is seldom used as an energy source but is involved in signaling pathways. Furthermore, treatment with exogenous gentiobiose induced budbreak in OWBs cultured in vitro, with increased concentrations of sulfur-containing amino acids, GSH, and ascorbate (AsA), as well as increased expression levels of the corresponding genes. Inhibition of GSH synthesis suppressed gentiobiose-induced budbreak accompanied by decreases in GSH and AsA concentrations and redox status. These results indicate that gentiobiose, a rare disaccharide, acts as a signal for dormancy release of gentian OWBs through the AsA-GSH cycle.  相似文献   

5.
A complex trisaccharide β-d-GalpNAcA-(1 → 4)-β-d-GlcpNAc-(1 → 4)-d-ManpNAc (3) was prepared in a good yield (35%) in a transglycosylation reaction catalyzed by β-N-acetylhexosaminidase from Talaromyces flavus using p-nitrophenyl 2-acetamido-2-deoxy-β-d-galacto-hexodialdo-1,5-pyranoside (1) as a donor followed by the in situ oxidation of the aldehyde functionality by NaClO2. The disaccharide β-d-GlcpNAc-(1 → 4)-d-ManpNAc (2) was used as galactosyl acceptor. A disaccharide β-d-GalpNAcA-(1 → 4)-d-GlcpNAc (4; 39%) originated as a by-product in the reaction. Oligosaccharides comprising a carboxy moiety at C-6 are shown to be very efficient ligands to natural killer cell activation receptors, particularly to human receptor CD69. Thus, oxidized trisaccharide 3 is the best-known oligosaccharidic ligand to this receptor, with IC50 = 2.5 × 10−9 M. The presented method of introducing a β-d-GalpNAcA moiety into carbohydrate structures is versatile and can be applied in the synthesis of other complex oligosaccharides.  相似文献   

6.
Enterobacterial common antigen (ECA) is expressed by Gram-negative bacteria belonging to Enterobacteriaceae, including emerging drug-resistant pathogens such as Escherichia coli, Klebsiella pneumoniae, and Proteus spp. Recent studies have indicated the importance of ECA for cell envelope integrity, flagellum expression, and resistance of enteric bacteria to acetic acid and bile salts. ECA, a heteropolysaccharide built from the trisaccharide repeating unit, →3)-α-d-Fucp4NAc-(1→4)-β-d-ManpNAcA-(1→4)-α-d-GlcpNAc-(1→, occurs as a cyclic form (ECACYC), a phosphatidylglycerol (PG)-linked form (ECAPG), and an endotoxin/lipopolysaccharide (LPS)-associated form (ECALPS). Since the discovery of ECA in 1962, the structures of ECAPG and ECACYC have been completely elucidated. However, no direct evidence has been presented to support a covalent linkage between ECA and LPS; only serological indications of co-association have been reported. This is paradoxical, given that ECA was first identified based on the capacity of immunogenic ECALPS to elicit antibodies cross-reactive with enterobacteria. Using a simple isolation protocol supported by serological tracking of ECA epitopes and NMR spectroscopy and mass spectrometry, we have succeeded in the first detection, isolation, and complete structural analysis of poly- and oligosaccharides of Shigella sonnei phase II ECALPS. ECALPS consists of the core oligosaccharide substituted with one to four repeating units of ECA at the position occupied by the O-antigen in the case of smooth S. sonnei phase I. These data represent the first structural evidence for the existence of ECALPS in the half-century since it was first discovered and provide insights that could prove helpful in further structural analyses and screening of ECALPS among Enterobacteriaceae species.  相似文献   

7.
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   

8.
O antigen (O polysaccharide) is an important and highly variable cell component present on the surface of cells which defines the serospecificity of Gram-negative bacteria. Most O antigens of Shigella flexneri, a cause of shigellosis, share a backbone composed of →2)-α-l-RhapIII-(1→2)-α-l-RhapII-(1→3)-α-l-RhapI-(1→3)-β-d-GlcpNAc-(1→ repeats, which can be modified by adding various substituents, giving rise to 19 serotypes. The known modifications include glucosylation on various sugar residues, O-acetylation on RhaI, and phosphorylation with phosphoethanolamine on RhaII or/and RhaIII. Recently, two new O-antigen modifications, namely, O-acetylation at position 3 or 4 of RhaIII and position 6 of GlcNAc, have been identified in several S. flexneri serotypes. In this work, the genetic basis for the 3/4-O-acetylation on RhaIII was elucidated. Bioinformatic analysis of the genome of S. flexneri serotype 2a strain Sf301, which carries 3/4-O-acetylation on RhaIII, revealed an O-acyltransferase gene designated oacB. Genetic studies combined with O-antigen structure analysis demonstrated that this gene is responsible for the 3/4-O-acetylation in serotypes 1a, 1b, 2a, 5a, and Y but not serotype 6, which has a different O-antigen backbone structure. The oacB gene is carried by a transposon-like structure located in the proA-adrA region on the chromosome, which represents a novel mechanism of mobilization of O-antigen modification factors in S. flexneri. These findings enhance our knowledge of S. flexneri O-antigen modifications and shed light on the origin of new O-antigen variants.  相似文献   

9.
The gram-positive bacterium Paenibacillus alvei CCM 2051T is covered by an oblique surface layer (S-layer) composed of glycoprotein subunits. The S-layer O-glycan is a polymer of [→3)-β-d-Galp-(1[α-d-Glcp-(1→6)]→4)-β-d-ManpNAc-(1→] repeating units that is linked by an adaptor of -[GroA-2→OPO2→4-β-d-ManpNAc-(1→4)]→3)-α-l-Rhap-(1→3)-α-l-Rhap-(1→3)-α-l-Rhap-(1→3)-β-d-Galp-(1→ to specific tyrosine residues of the S-layer protein. For elucidation of the mechanism governing S-layer glycan biosynthesis, a gene knockout system using bacterial mobile group II intron-mediated gene disruption was developed. The system is further based on the sgsE S-layer gene promoter of Geobacillus stearothermophilus NRS 2004/3a and on the Geobacillus-Bacillus-Escherichia coli shuttle vector pNW33N. As a target gene, wsfP, encoding a putative UDP-Gal:phosphoryl-polyprenol Gal-1-phosphate transferase, representing the predicted initiation enzyme of S-layer glycan biosynthesis, was disrupted. S-layer protein glycosylation was completely abolished in the insertional P. alvei CCM 2051T wsfP mutant, according to sodium dodecyl sulfate-polyacrylamide gel electrophoresis evidence and carbohydrate analysis. Glycosylation was fully restored by plasmid-based expression of wsfP in the glycan-deficient P. alvei mutant, confirming that WsfP initiates S-layer protein glycosylation. This is the first report on the successful genetic manipulation of bacterial S-layer protein glycosylation in vivo, including transformation of and heterologous gene expression and gene disruption in the model organism P. alvei CCM 2051T.Bacterial cell surface layer (S-layer) glycoproteins provide a unique self-assembly matrix that has been optimized by nature for regular and periodic display of glycans with nanometer scale accuracy (21, 31). Exploitation of this self-assembly system for surface display of functional, tailor-made glycans is an attractive alternative to the use of common cell surface anchors (7), with potential areas of application relating to any biological phenomenon that is based on carbohydrate recognition, such as receptor-substrate interaction, signaling, or cell-cell communication. A prerequisite for this endeavor is the availability of an S-layer glycoprotein-covered bacterium that is amenable to genetic manipulation. Despite the high application potential offered by the S-layer glycan display system, there are so far only two reports in the literature dealing with the genetic manipulation of S-layer glycoprotein-carrying bacteria. Both reports concern the gram-negative periodontal pathogen Tannerella forsythia ATCC 43037, but neither of them affects S-layer protein glycosylation (12, 24). In archaea, in contrast, molecular studies of S-layer protein glycosylation are quite advanced (1), but with the archaeal system, S-layer glycoprotein self-assembly, which is a prerequisite for the desired glycan display, has not been manageable in vitro so far.Our model organisms and, hence, candidates for S-layer-mediated glycan display enabled by carbohydrate engineering techniques are members of the Bacillaceae family. Currently, the S-layer glycosylation system of the thermophilic bacterium Geobacillus stearothermophilus NRS 2004/3a is best understood (20, 23, 29, 33, 34). However, a drawback of this organism is its resistance to take up foreign DNA. Although described in the literature (13, 14, 37), transformation of thermophilic bacilli seems to be a strain-specific trait. Based on successful transformation experiments in our laboratory, the mesophilic bacterium Paenibacillus alvei CCM 2051T (ATCC 6344; DSM 29) (formerly Bacillus alvei [4]) was chosen to set up a system for genetic manipulation. The bacterium is completely covered with an oblique S-layer lattice composed of glycoprotein species. Various aspects of its S-layer, including ultrastructural characterization (27), glycosylation analysis (2, 18), and glycan biosynthesis (11), have been investigated so far. The S-layer O-glycans are polymers of [→3)-β-d-Galp-(1[α-d-Glcp-(1→6)]→4)-β-d-ManpNAc-(1→] repeating units that are linked via the adaptor -[GroA-2→OPO2→4-β-d-ManpNAc-(1→4)]→3)-α-l-Rhap-(1→3)-α-l-Rhap-(1→3)-α-l-Rhap-(1→3)-β-d- Galp-(1→ to specific tyrosine residues (2, 18) of the S-layer protein SpaA (GenBank accession number FJ751775).Due to the presence of an identical adaptor saccharide backbone in the S-layer glycan of G. stearothermophilus NRS 2004/3a (29), where its biosynthesis is initiated by the UDP-Gal:phosphoryl-polyprenol Gal-1-phosphate transferase WsaP (33), it was conceivable that a homologous enzyme would initiate S-layer glycosylation in P. alvei CCM 2051T. Considering that the S-layer protein glycosylation machinery has been found to be encoded by S-layer glycosylation (slg) gene clusters (21), degenerate primers for the rml genes catalyzing the dTDP-l-Rha biosynthesis required for building up the adaptor saccharide of the P. alvei CCM 2051T S-layer glycan were used to define a point of entry into the glycosylation locus (K. Zarschler, B. Janesch, P. Messner, and C. Schäffer, unpublished data). Chromosome walking revealed the existence of an slg gene cluster of about 24 kb, including an open reading frame (ORF) predicted to encode the initiation enzyme of S-layer protein glycosylation. The corresponding gene, named wsfP, served as a first target for the gene knockout system developed in the course of the present study. This target was chosen because loss of function would be easily screenable, resulting in an S-layer glycosylation-deficient mutant. The gene knockout system constructed for insertional inactivation of the chromosomal wsfP gene of P. alvei CCM 2051T is based on the commercially available bacterial mobile group II intron Ll.LtrB of Lactococcus lactis, in combination with further components available in our laboratory, including the broad-host-range S-layer gene promoter of sgsE from G. stearothermophilus NRS 2004/3a (22) and the Geobacillus-Bacillus-Escherichia coli shuttle vector pNW33N. Bacterial mobile group II introns are retroelements inserted into specific DNA target sites at high frequency by use of the retrohoming mechanism, by which the excised intron lariat RNA is inserted directly into a DNA target site and is then reverse transcribed by the associated intron-encoded enzyme protein (6, 8, 17). Since the DNA target site is recognized primarily by base pairing of intron RNA, which can be modified, and a few intron-encoded-enzyme-protein recognition positions, these introns can be inserted efficiently into any specific DNA target (9, 15, 35, 40).The aim of this study is the development of a genetic tool for manipulation of S-layer protein glycosylation in P. alvei CCM 2051T. For proof of concept, we specifically deal with (i) the construction of a broad-host-range gene knockout system based on the L. lactis Ll.LtrB intron; (ii) its modification for specific disruption of the wsfP gene on the P. alvei CCM 2051T chromosome, encoding the putative initiation enzyme of S-layer glycan biosynthesis; and (iii) the reconstitution of enzyme activity by plasmid-based expression of wsfP and its predicted functional homologue wsaP from G. stearothermophilus NRS 2004/3a.  相似文献   

10.
An endoglucanase was isolated from cell walls of Zea mays seedlings. Characterization of the hydrolytic activity of this glucanase using model substrates indicated a high specificity for molecules containing intramolecular (1→3),(1→4)-β-d-glucosyl sequences. Substrates with (1→4)-β-glucosyl linkages, such as carboxymethylcellulose and xyloglucan were, degraded to a limited extent by the enzyme, whereas (1→3)-β-glucans such as laminarin were not hydrolyzed. When (1→3),(1→4)-β-d-glucan from Avena endosperm was used as a model substrate a rapid decrease in vicosity was observed concomitant with the formation of a glucosyl polymer (molecular weight of 1-1.5 × 104). Activity against a water soluble (1→3),(1→4)-β-d-glucan extracted from Zea seedling cell walls revealed the same depolymerization pattern. The size of the limit products would indicate that a unique recognition site exists at regular intervals within the (1→3),(1→4)-β-d-glucan molecule. Unique oligosaccharides isolated from the Zea (1→3),(1→4)-β-d-glucan that contained blocks of (1→4) linkages and/or more than a single contiguous (1→3) linkage were hydrolyzed by the endoglucanase. The unique regions of the (1→3),(1→4)-β-d-glucan may be the recognition-hydrolytic site of the Zea endoglucanase.  相似文献   

11.
Erwinia chrysanthemi (Dickeya dadantii) is a plant pathogenic bacterium that has a large capacity to degrade the plant cell wall polysaccharides. The present study reports the metabolic pathways used by E. chrysanthemi to assimilate the oligosaccharides sucrose and raffinose, which are particularly abundant plant sugars. E. chrysanthemi is able to use sucrose, raffinose, or melibiose as a sole carbon source for growth. The two gene clusters scrKYABR and rafRBA are necessary for their catabolism. The phenotypic analysis of scr and raf mutants revealed cross-links between the assimilation pathways of these oligosaccharides. Sucrose catabolism is mediated by the genes scrKYAB. While the raf cluster is sufficient to catabolize melibiose, it is incomplete for raffinose catabolism, which needs two additional steps that are provided by scrY and scrB. The scr and raf clusters are controlled by specific repressors, ScrR and RafR, respectively. Both clusters are controlled by the global activator of carbohydrate catabolism, the cyclic AMP receptor protein (CRP). E. chrysanthemi growth with lactose is possible only for mutants with a derepressed nonspecific lactose transport system, which was identified as RafB. RafR inactivation allows the bacteria to the assimilate the novel substrates lactose, lactulose, stachyose, and melibionic acid. The raf genes also are involved in the assimilation of α- and β-methyl-d-galactosides. Mutations in the raf or scr genes did not significantly affect E. chrysanthemi virulence. This could be explained by the large variety of carbon sources available in the plant tissue macerated by E. chrysanthemi.Pectinolytic erwiniae are enterobacteria that cause disease in a wide range of plants, including many crops of economic importance (23). The soft-rot symptom produced by Erwinia chrysanthemi (syn. Dickeya dadantii) results from the degradation of polysaccharides involved in the cohesion of the plant cell wall. The plant tissue maceration is concomitant with a large increase in the bacterial population (13). To ensure this multiplication, the bacteria assimilate various oligosaccharides released in the macerated tissue, which provide carbon and energy sources.E. chrysanthemi is known to use several carbon sources for growth, including sugars ranging from monosaccharides to polysaccharides. The completion of the E. chrysanthemi strain 3937 genome provides a genome-scale view into its potential catabolic capacities. A substantial part of the E. chrysanthemi genome is dedicated to genes involved in carbohydrate catabolism. In plant tissues, the most abundant soluble carbohydrates are the two oligosaccharides sucrose and raffinose (32). The trisaccharide raffinose [α-d-Galp-(1→6)-α-d-Glcp-(1⇆2)β-d-Fruf] and the related disaccharides sucrose [α-d-Glcp-(1⇆2)β-d-Fruf] and melibiose [α-d-Galp-(1→6)-d-Glcp] are used as carbon sources for E. chrysanthemi growth. Previous studies suggested links between the transport of lactose and that of raffinose and melibiose (15). The E. chrysanthemi wild-type strain 3937 does not use lactose [β-d-Galp-(1→4)-d-Glcp] as a carbon source for growth. This is due to the lack of a specific lactose transport system. However, spontaneous mutants able to assimilate lactose (designated Lac+) are easily obtained; they show a deregulation of the transport system LmrT, which is able to mediate lactose, melibiose, and raffinose transport (15). Despite our current knowledge of the strain 3937 genome sequence, no open reading frame (ORF) could be assigned to the lmrT gene, the identity of which remains unknown. We analyzed the E. chrysanthemi genome for the presence of potential genes involved in the catabolism of α-galactosides or α-glucosides. It contains a complete scrKYABR gene cluster that is involved in sucrose catabolism in various enterobacteria and a truncated rafRBA locus that is involved in raffinose catabolism. The growth with raffinose, despite the presence of an incomplete raf cluster, suggests that the missing functions are provided by other genes. Moreover, while E. chrysanthemi can catabolize melibiose, its genome does not contain homologues of the Escherichia coli melABR genes (30). Thus, to assimilate melibiose, E. chrysanthemi exploits other genes, which have yet to be identified. The present study mainly reports the role of the E. chrysanthemi gene clusters scr and raf in the catabolism of the oligosaccharides sucrose, raffinose, melibiose, and lactose. The importance of such catabolic pathways for bacterial multiplication in the plant tissues also was assessed during the infection process.  相似文献   

12.
Shigella flexneri O-antigen is an important and highly variable cell component presented on the outer leaflet of the outer membrane. Most Shigella flexneri bacteria share an O-antigen backbone composed of →2)-α-l-RhapIII-(1→2)-α-l-RhapII-(1→3)-α-l-RhapI-(1→3)-β-d-GlcpNAc-(1→ repeats, which can be modified by adding various chemical groups to different sugars, giving rise to diverse O-antigen structures and, correspondingly, to various serotypes. The known modifications include glucosylation on various sugar residues, O-acetylation on RhaI or/and RhaIII, and phosphorylation with phosphoethanolamine on RhaII or/and RhaIII. Recently, a new O-antigen modification, namely, O-acetylation at position 6 of N-acetylglucosamine (GlcNAc), has been identified in S. flexneri serotypes 2a, 3a, Y, and Yv. In this study, the genetic basis of the 6-O-acetylation of GlcNAc in S. flexneri was elucidated. An O-acyltransferase gene designated oacD was found to be responsible for this modification. The oacD gene is carried on serotype-converting bacteriophage SfII, which is integrated into the host chromosome by lysogeny to form a prophage responsible for the evolvement of serotype 2 of S. flexneri. The OacD-mediated 6-O-acetylation also occurs in some other S. flexneri serotypes that carry a cryptic SfII prophage with a dysfunctional gtr locus for type II glucosylation. The 6-O-acetylation on GlcNAc confers to the host a novel O-antigen epitope, provisionally named O-factor 10. These findings enhance our understanding of the mechanisms of the O-antigen variation and enable further studies to understand the contribution of the O-acetylation to the antigenicity and pathogenicity of S. flexneri.  相似文献   

13.
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   

14.
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   

15.
To metabolize both dietary fiber constituent carbohydrates and host glycans lining the intestinal epithelium, gut bacteria produce a wide range of carbohydrate-active enzymes, of which glycoside hydrolases are the main components. In this study, we describe the ability of phosphorylases to participate in the breakdown of human N-glycans, from an analysis of the substrate specificity of UhgbMP, a mannoside phosphorylase of the GH130 protein family discovered by functional metagenomics. UhgbMP is found to phosphorolyze β-d-Manp-1,4-β-d-GlcpNAc-1,4-d-GlcpNAc and is also a highly efficient enzyme to catalyze the synthesis of this precious N-glycan core oligosaccharide by reverse phosphorolysis. Analysis of sequence conservation within family GH130, mapped on a three-dimensional model of UhgbMP and supported by site-directed mutagenesis results, revealed two GH130 subfamilies and allowed the identification of key residues responsible for catalysis and substrate specificity. The analysis of the genomic context of 65 known GH130 sequences belonging to human gut bacteria indicates that the enzymes of the GH130_1 subfamily would be involved in mannan catabolism, whereas the enzymes belonging to the GH130_2 subfamily would rather work in synergy with glycoside hydrolases of the GH92 and GH18 families in the breakdown of N-glycans. The use of GH130 inhibitors as therapeutic agents or functional foods could thus be considered as an innovative strategy to inhibit N-glycan degradation, with the ultimate goal of protecting, or restoring, the epithelial barrier.  相似文献   

16.
Polyclonal antibodies raised against barley (1→3,1→4)-β-d-glucanase, α-amylase and carboxypeptidase were used to detect precursor polypeptides of these hydrolytic enzymes among the in vitro translation products of mRNA isolated from the scutellum and aleurone of germinating barley. In the scutellum, mRNA encoding carboxypeptidase appeared to be relatively more abundant than that encoding α-amylase or (1→3,1→4)-β-d-glucanase, while in the aleurone α-amylase and (1→3,1→4)-β-d-glucanase mRNAs predominated. The apparent molecular weights of the precursors for (1→3,1→4)-β-d-glucanase, α-amylase, and carboxypeptidase were 33,000, 44,000, and 35,000, respectively. In each case these are slightly higher (1,500-5,000) than molecular weights of the mature enzymes. Molecular weights of precursors immunoprecipitated from aleurone and scutellum mRNA translation products were identical for each enzyme.  相似文献   

17.
α-l-Arabinofuranosidases I and II were purified from the culture filtrate of Aspergillus awamori IFO 4033 and had molecular weights of 81,000 and 62,000 and pIs of 3.3 and 3.6, respectively. Both enzymes had an optimum pH of 4.0 and an optimum temperature of 60°C and exhibited stability at pH values from 3 to 7 and at temperatures up to 60°C. The enzymes released arabinose from p-nitrophenyl-α-l-arabinofuranoside, O-α-l-arabinofuranosyl-(1→3)-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose, and arabinose-containing polysaccharides but not from O-β-d-xylopyranosyl-(1→2)-O-α-l-arabinofuranosyl-(1→3)-O-β-d-xylopyranosyl-(1→4)-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose. α-l-Arabinofuranosidase I also released arabinose from O-β-d-xylopy-ranosyl-(1→4)-[O-α-l-arabinofuranosyl-(1→3)]-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose. However, α-l-arabinofuranosidase II did not readily catalyze this hydrolysis reaction. α-l-Arabinofuranosidase I hydrolyzed all linkages that can occur between two α-l-arabinofuranosyl residues in the following order: (1→5) linkage > (1→3) linkage > (1→2) linkage. α-l-Arabinofuranosidase II hydrolyzed the linkages in the following order: (1→5) linkage > (1→2) linkage > (1→3) linkage. α-l-Arabinofuranosidase I preferentially hydrolyzed the (1→5) linkage of branched arabinotrisaccharide. On the other hand, α-l-arabinofuranosidase II preferentially hydrolyzed the (1→3) linkage in the same substrate. α-l-Arabinofuranosidase I released arabinose from the nonreducing terminus of arabinan, whereas α-l-arabinofuranosidase II preferentially hydrolyzed the arabinosyl side chain linkage of arabinan.Recently, it has been proven that l-arabinose selectively inhibits intestinal sucrase in a noncompetitive manner and reduces the glycemic response after sucrose ingestion in animals (33). Based on this observation, l-arabinose can be used as a physiologically functional sugar that inhibits sucrose digestion. Effective l-arabinose production is therefore important in the food industry. l-Arabinosyl residues are widely distributed in hemicelluloses, such as arabinan, arabinoxylan, gum arabic, and arabinogalactan, and the α-l-arabinofuranosidases (α-l-AFases) (EC 3.2.1.55) have proven to be essential tools for enzymatic degradation of hemicelluloses and structural studies of these compounds.α-l-AFases have been classified into two families of glycanases (families 51 and 54) on the basis of amino acid sequence similarities (11). The two families of α-l-AFases also differ in substrate specificity for arabinose-containing polysaccharides. Beldman et al. summarized the α-l-AFase classification based on substrate specificities (3). One group contains the Arafur A (family 51) enzymes, which exhibit very little or no activity with arabinose-containing polysaccharides. The other group contains the Arafur B (family 54) enzymes, which cleave arabinosyl side chains from polymers. However, this classification is too broad to define the substrate specificities of α-l-AFases. There have been many studies of the α-l-AFases (3, 12), especially the α-l-AFases of Aspergillus species (28, 1215, 17, 22, 23, 2832, 3639, 4143, 46). However, there have been only a few studies of the precise specificities of these α-l-AFases. In previous work, we elucidated the substrate specificities of α-l-AFases from Aspergillus niger 5-16 (17) and Bacillus subtilis 3-6 (16, 18), which should be classified in the Arafur A group and exhibit activity with arabinoxylooligosaccharides, synthetic methyl 2-O-, 3-O-, and 5-O-arabinofuranosyl-α-l-arabinofuranosides (arabinofuranobiosides) (20), and methyl 3,5-di-O-α-l-arabinofuranosyl-α-l-arabinofuranoside (arabinofuranotrioside) (19).In the present work, we purified two α-l-AFases from a culture filtrate of Aspergillus awamori IFO 4033 and determined the substrate specificities of these α-l-AFases by using arabinose-containing polysaccharides and the core oligosaccharides of arabinoxylan and arabinan.  相似文献   

18.
The O antigen is the outer part of the lipopolysaccharide (LPS) in the outer membrane of Gram-negative bacteria and contains many repeats of an oligosaccharide unit. It contributes to antigenic variability and is essential to the full function and virulence of bacteria. Shigella is a Gram-negative human pathogen that causes diarrhea in humans. The O antigen of Shigella boydii type 14 consists of repeating oligosaccharide units with the structure [→6-d-Galpα1→4-d-GlcpAβ1→6-d-Galpβ1→4-d-Galpβ1→4-d-GlcpNAcβ1→]n. The wfeD gene in the O-antigen gene cluster of Shigella boydii type 14 was proposed to encode a galactosyltransferase (GalT) involved in O-antigen synthesis. We confirmed here that the wfeD gene product is a β4-GalT that synthesizes the Galβ1-4GlcNAcα-R linkage. WfeD was expressed in Escherichia coli, and the activity was characterized by using UDP-[3H]Gal as the donor substrate as well as the synthetic acceptor substrate GlcNAcα-pyrophosphate-(CH2)11-O-phenyl. The enzyme product was analyzed by liquid chromatography-mass spectrometry (LC-MS), high-performance liquid chromatography (HPLC), nuclear magnetic resonance (NMR), and galactosidase digestion. The enzyme was shown to be specific for the UDP-Gal donor substrate and required pyrophosphate in the acceptor substrate. Divalent metal ions such as Mn2+, Ni2+, and, surprisingly, also Pb2+ enhanced the enzyme activity. Mutational analysis showed that the Glu101 residue within a DxD motif is essential for activity, possibly by forming the catalytic nucleophile. The Lys211 residue was also shown to be required for activity and may be involved in the binding of the negatively charged acceptor substrate. Our study revealed that the β4-GalT WfeD is a novel enzyme that has virtually no sequence similarity to mammalian β4-GalT, although it catalyzes a similar reaction.Lipopolysaccharides (LPSs) consist of O-polysaccharide (O-antigenic) side chains covalently linked to a core polysaccharide and lipid A. LPSs are found in the outer membranes of Gram-negative bacteria, where they contribute to the structural integrity of the membrane and interact with the external environment (9, 10, 15). In the complex and dynamic microbial ecosystem of the human intestine, the communication between microorganisms and the gastrointestinal (GI) epithelium involves O-antigen and LPS binding molecules. Thus, the elimination of the O antigen may reduce virulence (2, 16, 21). Shigella is a genus of highly adapted bacterial pathogens that cause gastrointestinal disease, such as bacillary dysentery or shigellosis. A recent survey showed that shigellosis causes approximately 165 million cases of severe dysentery and more than 1 million deaths per year, mostly in children from developing countries (10). Shigella strains are categorized into four groups: S. boydii, S. dysenteriae, S. flexneri, and S. sonnei, each containing multiple subgroups of different serotypes, based on structural variations in their O antigens.O antigens consist of repeating units of oligosaccharides that are assembled individually, followed by the polymerization of units to form O antigens of different lengths. The glycosyltransferases involved in the biosynthesis of O antigens play a critical role in determining O-antigen structural diversity. The pentasaccharide repeating unit of S. boydii type 14 (B14) has the structure [→6-d-Galpα1→4-d-GlcpAβ1→6-d-Galpβ1→4-d-Galpβ1→4-d-GlcpNAcβ1→]n (12), suggesting the existence of five specific glycosyltransferases: a GlcNAc-phosphotransferase (WecA), three Gal-transferases, and a glucuronosyltransferase.Three distinct processes for the synthesis and translocation of O antigens have been described: the Wzx/Wzy-dependent pathway, the ATP binding cassette transporter-dependent process, and the synthase-dependent process (20, 25, 26). The biosynthesis of the S. boydii B14 O antigen that contains a variety of different sugar residues is expected to utilize the Wzy/Wzx-dependent pathway, where the synthesis of the repeating unit is initiated by WecA, catalyzing the transfer of sugar-phosphate (GlcNAcα-phosphate) from nucleotide sugar (UDP-GlcNAc) to a lipid carrier, undecaprenol-phosphate (Und-P), at the cytoplasmic side of the inner membrane. The wecA gene is present in the S. boydii B14 genome but outside the O-antigen gene cluster (1). The wecA gene is also involved in the synthesis of bacterial polysaccharides other than the O antigen. The extension of the chain is then mediated by specific glycosyltransferases that utilize nucleotide sugar donor substrates and are thought to be loosely associated with the inner membrane. In contrast, mammalian glycosyltransferases are usually membrane-bound proteins. Bacterial and mammalian glycosyltransferases, although they may have similar substrate specificities and form the same linkage, show significantly different amino acid sequences (4). Completed repeating units are then flipped across the membrane to the periplasmic side (by the flippase Wzx) and are polymerized (by Wzy) to form the O antigen under the control of a chain length regulator (Wzz). The repeating units are initially linked to the lipid carrier through GlcNAcα-phosphate. However, the S. boydii B14 O antigen has GlcNAc in the β linkage; thus, upon the polymerization of the completed repeating units, the linkage may be inverted, probably through the specific action of the polymerase Wzy. The entire polymer is then ligated to an outer core sugar based on lipid A. Upon completion, the LPS is extruded from the inner membrane and translocated to the outer membrane (19, 26). The latter-acting enzymes have multiple transmembrane regions that integrate them into the bacterial membranes.Genes involved in O-antigen biosynthesis are normally clustered between galF and gnd in Escherichia coli and Shigella and are classified into three different groups: (i) nucleotide sugar synthesis genes involved in the synthesis of donor substrates, (ii) glycosyltransferase genes, and (iii) O-antigen-processing genes, such as the flippase gene wzx and the polymerase gene wzy. The O-antigen gene cluster of B14 has been sequenced and analyzed (10). Four putative glycosyltransferase genes found in the B14 O-antigen synthesis gene cluster are wfeA, wfeB, wfeD, and wfeE. WfeD shares 38% identity and 57% similarity to the putative glycosyltransferase Orf9, which is involved in the synthesis of the E. coli O136 O antigen (our unpublished data). Since the O antigens of B14 and E. coli O136 share only one common linkage, d-Galpβ1→4-d-GlcpNAc (12, 23), wfeD was proposed to encode the galactosyltransferase (GalT) that transfers Gal to GlcNAcα-PP-Und in the β1-4 linkage, which is the second step in the biosynthetic pathway of the B14 O-antigen repeating unit.We have used biochemical approaches to assay the WfeD enzyme activity and to characterize this enzyme. The lipid carrier analog GlcNAcα-PO3-PO3-(CH2)11-O-phenyl [GlcNAc-PP-(CH2)11-OPh] has previously been used as a defined synthetic acceptor substrate for the characterization of glycosyltransferases from E. coli serotypes O7 (β1,3-GalT WbbD), O56 (β1,3-Glc-transferase WfaP), and O152 (β1,3-Glc-transferase WfgD) (6, 17). In this work, we showed that GlcNAc-PP-(CH2)11-OPh could also serve as an exogenous substrate for WfeD from B14. We were therefore able to prove that wfeD encodes a novel β1,4-GalT.  相似文献   

19.
A large panel of fungal β-N-acetylhexosaminidases was tested for the regioselectivity of the β-GlcNAc transfer onto galacto-type acceptors ( -galactose, lactose, 2-acetamido-2-deoxy- -galactopyranose). A unique, non-reducing disaccharide β- -GlcpNAc-(1→1)-β- -Galp and trisaccharides β- -GlcpNAc-(1→4)-β- -GlcpNAc-(1→1)-β- -Galp, β- -Galp-(1→4)-β- -Glcp-(1→1)-β- -GlcpNAc and β- -Galp-(1→4)-α- -Glcp-(1→1)-β- -GlcpNAc were synthesised under the catalysis of the β-N-acetylhexosaminidase from the Aspergillus flavofurcatis CCF 3061 with -galactose and lactose as acceptors. The use of 2-acetamido-2-deoxy- -galactopyranose as an acceptor with the β-N-acetylhexosaminidases from A. flavofurcatis CCF 3061, A. oryzae CCF 1066 and A. tamarii CCF 1665 afforded only β- -GlcpNAc-(1→6)- -GalpNAc.  相似文献   

20.
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   

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