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 625 毫秒
1.
Hemicelluloses of cell walls of a proso millet cell suspension culture   总被引:1,自引:1,他引:0  
Cell wall composition of a stable suspension of proso millet (Panicum miliaceum L. cv Abarr) cells is similar to those of tissues and cell suspensions of other graminaceous species. Extraction of hemicelluloses with step-wise increasing concentrations of alkali yields materials that, like those of embryonal cells of maize coleoptiles, comprise mostly glucuronoarabinoxylan, xyloglucan, and small amounts of (1-3),(1-4)-β-d-glucan. As in the walls of embryonal cells of the maize coleoptile, 5-arabinosyl and 3-arabinosyl comprise much higher proportions of the total hemicellulosic sugars than in walls of developed or elongated cells. Unlike cells of many dicotyledonous species, millet cells do not elongate or undergo observable differentiation during the stationary phase of culture, and consequently, their wall composition is remarkably consistent throughout the culture cycle. The proso millet cell suspension culture constitutes a reasonable model for study of cell wall biogenesis in embryonal cells of a graminaceous species, but because of marked changes in the composition of hemicelluloses in these species during cell enlargement, additional model systems should be sought.  相似文献   

2.
The composition of the cell wall of Fusicoccum amygdali   总被引:1,自引:1,他引:0       下载免费PDF全文
1. The cell wall of Fusicoccum amygdali consisted of polysaccharides (85%), protein (4–6%), lipid (5%) and phosphorus (0.1%). 2. The main carbohydrate constituent was d-glucose; smaller amounts of d-glucosamine, d-galactose, d-mannose, l-rhamnose, xylose and arabinose were also identified, and 16 common amino acids were detected. 3. Chitin, which accounted for most of the cell-wall glucosamine, was isolated in an undegraded form by an enzymic method. Chitosan was not detected, but traces of glucosamine were found in alkali-soluble and water-soluble fractions. 4. Cell walls were stained dark blue by iodine and were attacked by α-amylase, with liberation of glucose, maltose and maltotriose, indicating the existence of chains of α-(1→4)-linked glucopyranose residues. 5. Glucose and gentiobiose were liberated from cell walls by the action of an exo-β-(1→3)-glucanase, giving evidence for both β-(1→3)- and β-(1→6)-glucopyranose linkages. 6. Incubation of cell walls with Helix pomatia digestive enzymes released glucose, N-acetyl-d-glucosamine and a non-diffusible fraction, containing most of the cell-wall galactose, mannose and rhamnose. Part of this fraction was released by incubating cell walls with Pronase; acid hydrolysis yielded galactose 6-phosphate and small amounts of mannose 6-phosphate and glucose 6-phosphate as well as other materials. Extracellular polysaccharides of a similar nature were isolated and may be formed by the action of lytic enzymes on the cell wall. 7. About 30% of the cell wall was resistant to the action of the H. pomatia digestive enzymes; the resistant fraction was shown to be a predominantly α-(1→3)-glucan. 8. Fractionation of the cell-wall complex with 1m-sodium hydroxide gave three principal glucan fractions: fraction BB had [α]D +236° (in 1m-sodium hydroxide) and showed two components on sedimentation analysis; fraction AA2 had [α]D −71° (in 1m-sodium hydroxide) and contained predominantly β-linkages; fraction AA1 had [α]D +40° (in 1m-sodium hydroxide) and may contain both α- and β-linkages.  相似文献   

3.
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   

4.
Hoson T  Masuda Y  Nevins DJ 《Plant physiology》1992,98(4):1298-1303
Polyclonal antibodies, raised against β-d-glucans prepared from oat (Avena sativa L.) caryopses, cross-reacted specifically with (1→3),(1→4)-β-d-glucans when challenged in a dot blot analysis of related polymers bound to a cellulose thin layer chromatography plate. The antibodies suppressed indoleacetic acid (IAA)-induced elongation of segments from maize (Zea mays L.) coleoptiles when the outer surface was abraded. However, IAA-induced elongation of nonabraded segments or segments with abrasion restricted to the interior of the cylinder was not influenced by the antibodies. Fab fragments prepared from the antibodies gave similar results. The capacity for IAA to overcome outward curvature of split coleoptile segments was partially reversed by treatment of the segments with the antibodies. Fluorescence microscopy revealed that antibody penetration was largely restricted to the epidermal cell wall region. These results support the view that the degradation of (1→3),(1→4)-β-d-glucans in the outer epidermal cell wall serves an essential role in auxin-induced elongation of Poaceae coleoptiles.  相似文献   

5.
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   

6.
Three-day-cultured cells of Vinca rosea L. (in the cell division phase) and 5-day-cultured cells (in the cell expansion phase) prelabelled with d -[U-14C] glucose were incubated in a medium containing unlabelled glucose. After various periods of chase, extra-cellular polysaccharides (ECP) and cell walls were isolated, and cell walls were fractionated into pectic substances, hemicellulose, and cellulose fractions. After acid hydrolysis, the radioactive constituents in the pectic substances and hemicellulose fractions were analyzed. Active turnover was observed in arabinose and galactose in the hemicellulose fraction of cell walls, while the constituents of the pectic substances, and xylose and glucose in the hemicellulose fraction did not undergo active turnover. The proportion of radioactivities of arabinose and galactose in total radioactivity of ECP increased markedly after chasing. These results indicate that arabinogalactan was synthesized, deposited in the cell wall, degraded rapidly, and made soluble in the medium as a part of ECP.  相似文献   

7.
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   

8.
Excised Zea mays L. embryos were cultured on Linsmaier and Skoog medium. Coleoptiles were sampled at regular intervals and the length, fresh weight, cell wall weight, and cell wall neutral sugar composition were determined. A specific β-d-glucanase from Bacillus subtilis was used to determine the content of a (1 → 3),(1 → 4)-β-d-glucan.  相似文献   

9.
The complement system is part of our first line of defense against invading pathogens. The strategies used by Enterococcus faecalis to evade recognition by human complement are incompletely understood. In this study, we identified an insertional mutant of the wall teichoic acid (WTA) synthesis gene tagB in E. faecalis V583 that exhibited an increased susceptibility to complement-mediated killing by neutrophils. Further analysis revealed that increased killing of the mutant was due to a higher rate of phagocytosis by neutrophils, which correlated with higher C3b deposition on the bacterial surface. Our studies indicated that complement activation via the lectin pathway was much stronger on the tagB mutant compared with wild type. In concordance, we found an increased binding of the key lectin pathway components mannose-binding lectin and mannose-binding lectin-associated serine protease-2 (MASP-2) on the mutant. To understand the mechanism of lectin pathway inhibition by E. faecalis, we purified and characterized cell wall carbohydrates of E. faecalis wild type and V583ΔtagB. NMR analysis revealed that the mutant strain lacked two WTAs with a repeating unit of →6)[α-l-Rhap-(1→3)]β-d-GalpNAc-(1→5)-Rbo-1-P and →6) β-d-Glcp-(1→3) [α-d-Glcp-(1→4)]-β-d-GalpNAc-(1→5)-Rbo-1-P→, respectively (Rbo, ribitol). In addition, compositional changes in the enterococcal rhamnopolysaccharide were noticed. Our study indicates that in E. faecalis, modification of peptidoglycan by secondary cell wall polymers is critical to evade recognition by the complement system.  相似文献   

10.
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   

11.
Peptidoglycan hydrolases (PGHs) are responsible for bacterial cell lysis. Most PGHs have a modular structure comprising a catalytic domain and a cell wall-binding domain (CWBD). PGHs of bacteriophage origin, called endolysins, are involved in bacterial lysis at the end of the infection cycle. We have characterized two endolysins, Lc-Lys and Lc-Lys-2, identified in prophages present in the genome of Lactobacillus casei BL23. These two enzymes have different catalytic domains but similar putative C-terminal CWBDs. By analyzing purified peptidoglycan (PG) degradation products, we showed that Lc-Lys is an N-acetylmuramoyl-l-alanine amidase, whereas Lc-Lys-2 is a γ-d-glutamyl-l-lysyl endopeptidase. Remarkably, both lysins were able to lyse only Gram-positive bacterial strains that possess PG with d-Ala4d-Asx-l-Lys3 in their cross-bridge, such as Lactococcus casei, Lactococcus lactis, and Enterococcus faecium. By testing a panel of L. lactis cell wall mutants, we observed that Lc-Lys and Lc-Lys-2 were not able to lyse mutants with a modified PG cross-bridge, constituting d-Ala4l-Ala-(l-Ala/l-Ser)-l-Lys3; moreover, they do not lyse the L. lactis mutant containing only the nonamidated d-Asp cross-bridge, i.e. d-Ala4d-Asp-l-Lys3. In contrast, Lc-Lys could lyse the ampicillin-resistant E. faecium mutant with 3→3 l-Lys3-d-Asn-l-Lys3 bridges replacing the wild-type 4→3 d-Ala4-d-Asn-l-Lys3 bridges. We showed that the C-terminal CWBD of Lc-Lys binds PG containing mainly d-Asn but not PG with only the nonamidated d-Asp-containing cross-bridge, indicating that the CWBD confers to Lc-Lys its narrow specificity. In conclusion, the CWBD characterized in this study is a novel type of PG-binding domain targeting specifically the d-Asn interpeptide bridge of PG.  相似文献   

12.
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   

13.
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   

14.
Background and aims Cress-seed (Lepidium sativum) exudate exerts an allelochemical effect, promoting excessive hypocotyl elongation and inhibiting root growth in neighbouring Amaranthus caudatus seedlings. We investigated acidic disaccharides present in cress-seed exudate, testing the proposal that the allelochemical is an oligosaccharin—lepidimoic acid (LMA; 4-deoxy-β-l-threo-hex-4-enopyranuronosyl-(1→2)-l-rhamnose).Methods Cress-seed exudate was variously treated [heating, ethanolic precipitation, solvent partitioning, high-voltage paper electrophoresis and gel-permeation chromatography (GPC)], and the products were bioassayed for effects on dark-grown Amaranthus seedlings. Two acidic disaccharides, including LMA, were isolated and characterized by electrophoresis, thin-layer chromatography (TLC) and nuclear magnetic resonance (NMR) spectroscopy, and then bioassayed.Key Results Cress-seed exudate contained low-Mr, hydrophilic, heat-stable material that strongly promoted Amaranthus hypocotyl elongation and inhibited root growth, but that separated from LMA on electrophoresis and GPC. Cress-seed exudate contained ∼250 µm LMA, whose TLC and electrophoretic mobilities, susceptibility to mild acid hydrolysis and NMR spectra are reported. A second acidic disaccharide, present at ∼120 µm, was similarly characterized, and shown to be β-d-xylopyranosyl-(1→3)-d-galacturonic acid (Xyl→GalA), a repeat unit of xylogalacturonan. Purified LMA and Xyl→GalA when applied at 360 and 740 µm, respectively, only slightly promoted Amaranthus hypocotyl growth, but equally promoted root growth and thus had no effect on the hypocotyl:root ratio, unlike total cress-seed exudate.Conclusions LMA is present in cress seeds, probably formed by rhamnogalacturonan lyase action on rhamnogalacturonan-I during seed development. Our results contradict the hypothesis that LMA is a cress allelochemical that appreciably perturbs the growth of potentially competing seedlings. Since LMA and Xyl→GalA slightly promoted both hypocotyl and root elongation, their effect could be nutritional. We conclude that rhamnogalacturonan-I and xylogalacturonan (pectin domains) are not sources of oligosaccharins with allelochemical activity, and the biological roles (if any) of the disaccharides derived from them are unknown. The main allelochemical principle in cress-seed exudate remains to be identified.  相似文献   

15.
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   

16.
With the exception of cellulose and callose, the cell wall polysaccharides are synthesized in Golgi membranes, packaged into vesicles, and exported to the plasma membrane where they are integrated into the microfibrillar structure. Consistent with this paradigm, several published reports have shown that the maize (Zea mays) mixed-linkage (1→3),(1→4)-β-d-glucan, a polysaccharide that among angiosperms is unique to the grasses and related Poales species, is synthesized in vitro with isolated maize coleoptile Golgi membranes and the nucleotide-sugar substrate, UDP-glucose. However, a recent study reported the inability to detect the β-glucan immunocytochemically at the Golgi, resulting in a hypothesis that the mixed-linkage β-glucan oligomers may be initiated at the Golgi but are polymerized at the plasma membrane surface. Here, we demonstrate that (1→3),(1→4)-β-d-glucans are detected immunocytochemically at the Golgi of the developing maize coleoptiles. Further, when maize seedlings at the third-leaf stage were pulse labeled with [14C]O2 and Golgi membranes were isolated from elongating cells at the base of the developing leaves, (1→3),(1→4)-β-d-glucans of an average molecular mass of 250 kD and higher were detected in isolated Golgi membranes. When the pulse was followed by a chase period, the labeled polysaccharides of the Golgi membrane diminished with subsequent transfer to the cell wall. (1→3),(1→4)-β-d-Glucans of at least 250 kD were isolated from cell walls, but much larger aggregates were also detected, indicating a potential for intermolecular interactions with glucuronoarabinoxylans or intermolecular grafting in muro.An overwhelming body of evidence accumulated has established that the (1→4)-β-d-glucan chains of cellulose microfibrils are synthesized and assembled at the plasma membrane surface (Delmer, 1999; Saxena and Brown, 2005), whereas, with the lone exception of the (1→3)-β-d-glucan, callose, all noncellulosic pectin and cross-linking glycan polysaccharides are synthesized in Golgi membranes (Northcote and Pickett-Heaps, 1966; Ray et al., 1969, 1976; Harris and Northcote, 1971; Zhang and Staehelin, 1992). Using several plant systems, including grass species, autoradiography and membrane fractionation showed that monosaccharides from 14C-labeled substrates accumulated in cell wall polysaccharides in Golgi vesicles during a pulse were subsequently transferred to the cell wall when chased with unlabeled substrates (Northcote and Pickett-Heaps, 1966; Pickett-Heaps, 1967; Jilka et al., 1972). Early studies showed that labeled sugars from nucleotide-sugar substrates could be incorporated into alcohol-insoluble polysaccharides using microsomal membranes, and later refined by isolation of Golgi membranes and the synthesis of defined polysaccharides with combinations of nucleotide sugars (Bailey and Hassid, 1966; Ray et al., 1969, 1976; Smith and Stone, 1973; Ray, 1980; Hayashi and Matsuda, 1981a; Gordon and Maclachlan, 1989; Gibeaut and Carpita, 1993).When micromolar concentrations of substrates were used, only small chains of the glycan products were typically made in vitro. For example, xyloglucan oligomers with the characteristic α-d-Xyl-(1→6)-d-glucosyl unit, isoprimeverose, were synthesized with isolated microsomal membranes and low concentrations of UDP-Glc and UDP-Xyl (Ray et al., 1976; Hayashi and Matsuda, 1981b). When concentrations of each nucleotide sugar were increased to millimolar concentrations, then polysaccharides of about 250 kD were synthesized containing the characteristic XXXG heptasaccharide unit structure (Gordon and Maclachlan, 1989). Immunocytochemical evidence with antibodies directed against the terminal nonreducing xylosyl and fucosyl residues confirm that synthesis of the xyloglucan backbone begins in the cis-Golgi membrane and culminates with fucosylation in the trans-Golgi membrane and trans-Golgi network (Moore et al., 1991; Lynch and Staehelin, 1992; Zhang and Staehelin, 1992). The fucosyl transferase responsible for xyloglucan side chain decoration was also shown to be a Golgi-resident protein by in vitro synthesis of xyloglucan polymers (Camirand and Maclachlan, 1986).In Poales species, including all grasses, the mixed-linkage (1→3),(1→4)-β-d-glucan is a major cross-linking glycan that appears transiently during cell elongation in growing tissues and accumulates to large amounts in the cell walls of the endosperm of certain grains (Stone and Clarke, 1992; Trethewey et al., 2005). Bailey and Hassid (1966) demonstrated the synthesis in vitro of noncellulosic glucans with microsomal membranes from grasses. Henry and Stone (1982) used the Bacillus subtilis endoglucanase, an enzyme that generates diagnostic cellodextrin-(1→3)-β-Glc units from (1→3),(1→4)-β-d-glucan to show that the mixed-linkage β-glucan was made specifically with UDP-Glc and microsomal membranes. We used flotation centrifugation to obtain highly enriched Golgi membranes from which (1→3),(1→4)-β-d-glucans of an average of about 250 kD were synthesized (Gibeaut and Carpita, 1993).The BG1 monoclonal antibody recognizes the (1→3),(1→4)-β-d-glucan with high specificity (Meikle et al., 1994). This monoclonal antibody has been used to show dramatic changes in epitope abundance of (1→3),(1→4)-β-d-glucan in the cell walls of developing tissues (Meikle et al., 1994; Trethewey et al., 2005; McCann et al., 2007) and its appearance in the cell walls of Arabidopsis (Arabidopsis thaliana) following heterologous expression of genes thought to encode its synthases (Burton et al., 2006; Doblin et al., 2009). The failure to detect (1→3),(1→4)-β-d-glucan in Golgi membranes and only in the cell wall prompted Fincher (2009) to conclude that cellodextrin oligomers of the (1→3),(1→4)-β-d-glucan may be initiated in the Golgi membrane, but the actual polymerization of the polysaccharide occurs at the plasma membrane.While there is little question that synthesis of full-length polymers is possible in vitro with isolated Golgi membranes and UDP-Glc (Gibeaut and Carpita, 1993; Buckeridge et al., 1999, 2001; Urbanowicz et al., 2004), Fincher (2009) asserts correctly that there exists no experimental evidence that the polymer is made in vivo within the Golgi membrane in intact tissues. In fact, earlier work showing the paucity of immunolabeling of (1→3),(1→4)-β-d-glucan in Golgi membranes of developing wheat (Triticum aestivum) endosperm at a time of active deposition called to question the site of synthesis in vivo (Philippe et al., 2006). There is precedence for the synthesis of chitin in vitro with precociously activated chitisomes (Bracker et al., 1976), a vesicular package of chitin synthase that in vivo is quiescent until reaching the plasma membrane. No activity of chitin synthase from isolated plasma membranes could be demonstrated. In a similar way, the Golgi synthase activity of (1→3),(1→4)-β-d-glucan could be a precocious activation in vitro of a plasma membrane activity.As in vitro synthesis studies clearly show synthesis of full-length (1→3),(1→4)-β-d-glucan only at the Golgi, we reexamined the puzzling finding of its absence from Golgi bodies to determine the true site of synthesis in vivo. In contrast to Fincher (2009), our own immunocytochemistry shows (1→3),(1→4)-β-d-glucan is indeed in the Golgi membrane in 2-d-old coleoptiles, when rapid growth is just beginning. However, we are unable to detect the β-glucan in Golgi after the peak rate of elongation. We pulse labeled maize (Zea mays) seedlings with radiolabeled CO2 and followed the fate of label captured by photosynthesis and translocated to elongating cells at the base of the seedling. We found by flotation centrifugation that Golgi membranes contain (1→3),(1→4)-β-d-glucan of at least 250 kD, similar to that of the product of in vitro synthesis at optimal UDP-Glc concentrations and commercial preparations of barley (Hordeum vulgare) endosperm (1→3),(1→4)-β-d-glucan (Gibeaut and Carpita, 1993; Buckeridge et al., 1999, 2001; Urbanowicz et al., 2004). When polysaccharides are extracted sequentially from the cell walls by hot ammonium oxalate, and increasing concentrations of NaOH to 4 m, the (1→3),(1→4)-β-d-glucans are found mostly in the higher concentrations of alkali fractions. While 250 kD polymers are observed, most of the (1→3),(1→4)-β-d-glucans eluted in fractions containing glucuronoarabinoxylans (GAXs), which are much larger, indicating either that an aggregation with GAXs increase the apparent size or that trans-glucosylation events increase the degree of polymerization of the (1→3),(1→4)-β-d-glucans.  相似文献   

17.
Polyclonal antibodies were raised in rabbits in response to the administration of purified exo- and endoglucanases extracted from cell walls of maize (Zea mays L. B37 × Mo17) coleoptiles. Since the antibodies formed specific conjugates when challenged with the glucanase antigens in immunoblot assays they were employed to evaluate the participation of glucanases in tissue growth. Indole-3-acetic acid induced cell elongation of abraded coleoptile segments was inhibited when the antibodies were supplied as a short term pretreatment (25-200 microgram/milliliter of serum protein). The extent of inhibition of IAA induced cell elongation was additive when endo- and exoglucanase antibodies were applied together. The results suggest that both enzymes have a role in mediating IAA-induced cell elongation. Pretreatment with exo- and endoglucanases antibodies also inhibited IAA induced degradation of noncellulosic β-d-glucans and the increased level of cellulosic polymers in maize coleoptiles. Antibodies also inhibited the expression of the autohydrolytic degradation of glucans in isolated cell walls. The extent of inhibition was dependent on the antibody concentration applied. The results support the contention that enzymatic processes mediated by exo- and endoglucanases are responsible for cell wall autolytic reactions and that these reactions are linked to the mechanism for expressing auxin induced cell elongation in maize coleoptiles.  相似文献   

18.
Particulate, and digitonin-solubilized, enzyme systems from Phaseolus aureus and Lupinus albus catalyze the biosynthesis of aqueous-insoluble glucans from UDP-d-glucose. The digitonin treatment greatly increases the enzymic activity of (per unit protein) both the 34,000g pellet and the supernatant liquid as compared with that of the original particles. Most of the polymer produced (90-95%) is soluble in hot, dilute alkali; the interglucosidic linkages of the alkali-soluble and alkali-insoluble polymers are identical. The optimum concentration for the incorporation of radioactivity from UDP-d-glucose-14C into soluble glucan is high; at 10−3 m at least 50% of the added radioactive glucosyl donor is incorporated.  相似文献   

19.
Tsai CM  Hassid WZ 《Plant physiology》1973,51(6):998-1001
UDP-d-glucose, at a micromolar level in the presence of MgCl2 and oat (Avena sativa) coleoptile particulate enzyme which contains both β-(1 → 3) and β-(1 → 4) glucan synthetases, produces glucan with mainly β-(1 → 4) glucosyl linkages. An activation of β-(1 → 3) glucan synthetase by UDP-d-glucose and a decrease in the formation of β-(1 → 3) glucan in the presence of MgCl2 have been observed. However, at high substrate concentration (≥ 10−4m), the activation of β-(1 → 3) glucan synthetase is so pronounced that the formation of β-(1 → 3) glucosyl linkage predominates in synthesized glucan regardless of the presence of MgCl2. These observations may explain the striking shift in the composition of glucan of particulate enzyme from a β-(1 → 4) to β-(1 → 3) glucosyl linkage when UDP-d-glucose concentration is raised from a low concentration (≤ 10−5m) to a higher concentration (≥ 10−4m).  相似文献   

20.
1. When pig ear skin slices were cultured for 18h in the presence of 1μg of tunicamycin/ml the incorporation of d-[3H]glucosamine into the epidermis, solubilized with 8m-urea/5% (w/v) sodium dodecyl sulphate, was inhibited by 45–55%. This degree of inhibition was not increased by using up to 5μg of tunicamycin/ml or by treating the skin slices with tunicamycin for up to 8 days. The incorporation of (U-14C)-labelled l-amino acids under these conditions was not affected by tunicamycin. Polyacrylamide-gel electrophoresis indicated that the labelling of the major glycosaminoglycan peak with d-[3H]glucosamine was unaffected, whereas that of the faster migrating glycoprotein components was considerably decreased in the presence of tunicamycin. 2. Subcellular fractionation indicated that tunicamycin specifically inhibited the incorporation of d-[3H]glucosamine but not of (U-14C)-labelled l-amino acids into particulate (mainly plasma-membrane) glycoproteins by about 70%. The labelling of soluble glycoproteins was hardly affected. Polyacrylamide-gel electrophoresis of the plasma-membrane fraction showed decreased d-[3H]glucosamine incorporation into all glycoprotein components, indicating that the plasma-membrane glycoproteins contained mainly N-asparagine-linked oligosaccharides. 3. Cellulose acetate electrophoresis of both cellular and extracellular glycosaminoglycans showed that tunicamycin had no significant effect on the synthesis of the major component, hyaluronic acid. However, the incorporation of both d-[3H]glucosamine and 35SO42− into sulphated glycosaminoglycans was inhibited by about 50%. This inhibition was partially overcome, at least in the cellular fraction, by 2mm-p-nitrophenyl β-d-xyloside indicating that tunicamycin-treated epidermis retained the ability to synthesize sulphated glycosaminoglycan chains. Tunicamycin may affect the synthesis and/or degradation of proteoglycan core proteins or the xylosyltransferase. 4. Electron-microscopic examination of epidermis treated with tunicamycin for up to 4 days revealed no significant changes in cell-surface morphology or in epidermal-cell adhesion. Either N-asparagine-linked carbohydrates play little role in epidermal-cell adhesion or more probably there is little turnover of these components in epidermal adhesive structures such as desmosomes and hemidesmosomes during organ culture.  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