首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Glucansucrases are responsible for the production of α-glucans using sucrose as the substrate. Glucansucrases may also produce different oligosaccharides by transferring the glucose moiety from sucrose to a variety carbohydrates acting as acceptor nucleophile. In this study, melibiose-derived oligosaccharides were produced by the glucansucrase from Lactobacillus reuteri E81 expressed without the N-terminal region (GtfA-ΔN). The reaction products were characterized by TLC, LC-MS and NMR analysis and it was found that GtfA-ΔN synthesized melibiose-derived oligosaccharides (DP3 and DP4) by adding glucose units through alpha 1->3 or 1->6 glycosidic bond. The functional characteristics of these melibiose-derived oligosaccharides were determined by testing the immune-modulatory functions in HT-29 cells and testing their growth promoting effects for important probiotic and pathogenic strains. The melibiose-derived oligosaccharides triggered the production of anti-inflammatory cytokine IL-10 and pro-inflammatory cytokine TNF-α depending on their concentrations. Finally, melibiose-derived oligosaccharides showed bifidogenic effect as potential prebiotics.  相似文献   

2.
Cellulase reaction mechanism was investigated with the use of following pure cellulosic substrates: Microcrystalline cellulose (Avicel), α‐cellulose (Sigma), filter paper, cotton, and non‐crystalline cellulose (NCC). NCC is amorphous cellulose prepared in our laboratory by treatment with concentrated sulfuric acid. When hydrolyzed with cellulase, NCC produces significant amount of cello‐oligosaccharides (COS) as reaction intermediates along with glucose and cellobiose. The COS produced by cellulase were categorized into two different moieties based upon their degree of polymerization (DP): low DP (less than 7) COS (LD‐COS) and high DP COS (HD‐COS). Endo‐glucanase (Endo‐G) reacts rapidly on the NCC reducing its DP to 30–60, after which the Endo‐G reaction with NCC ceases. HD‐COS is produced from NCC by the action of Endo‐G, whereas LD‐COS is produced by exo‐glucanase (Exo‐G). β‐Glucosidase (β‐G) hydrolyzes LD‐COS to produce cellobiose, but it does not hydrolyze HD‐COS. DP of NCC affects the action of Exo‐G in such a way that the overall yield is high for high DP NCC. This is in line with previous findings that substrate‐recognition by Exo‐G requires binding on β‐glucan chain with DP of 10 for the hydrolysis to take place. The individual cellulose chain residues within solid having DP less than 10 therefore remain unreacted. The percentage of the unreacted portion would be lower for high DP NCC, which results high overall conversion. The surface area and the number of reactive sites on the substrate facilitate adsorption of enzyme therefore the initial rate of the hydrolysis. The overall extent of conversion of cellulose, however, is controlled primarily by its inherent characteristics such as DP and crystallinity. Biotechnol. Bioeng. 2009;102: 1570–1581. © 2008 Wiley Periodicals, Inc.  相似文献   

3.
4.
β-Glucosidases designated MoCel3A and MoCel3B were successfully overexpressed in Magnaporthe oryzae. MoCel3A and MoCel3B showed optimal activity at 50 °C and pH 5.0–5.5. MoCel3A exhibited higher activity on higher degree of polymerization (DP) oligosaccharides and on β-1,3-linked oligosaccharides than on β-1,4-linked oligosaccharides. Furthermore, MoCel3A could liberate glucose from polysaccharides such as laminarin, 1,3-1,4-β-glucan, phosphoric acid-swollen cellulose, and pustulan, of which laminarin was the most suitable substrate. Conversely, MoCel3B preferentially hydrolyzed lower DP oligosaccharides such as cellobiose, cellotriose, and laminaribiose. Furthermore, the synergistic effects of combining enzymes including MoCel3A and MoCel3B were investigated. Depolymerization of 1,3-1,4-β-glucan by M. oryzae cellobiohydrolase (MoCel6A) enhanced the production of glucose by the actions of MoCel3A and MoCel3B. In these reactions, MoCel3A hydrolyzed higher DP oligosaccharides, resulting in the release of glucose and cellobiose, and MoCel3B preferentially hydrolyzed lower DP oligosaccharides including cellobiose. On the other hand, MoCel3A alone produced glucose from laminarin at levels equivalent to 80% of maximal hydrolysis obtained by the combined action of MoCel3A, MoCel3B, and endo-1,3-β-glucanase. Therefore, MoCel3A and MoCel3B activities yield glucose from not only cellulosic materials but also hemicellulosic polysaccharides.  相似文献   

5.
Clostridium thermocellum cellodextrin phosphorylase (CtCDP), a single-module protein without an apparent carbohydrate-binding module, has reported activities on soluble cellodextrin with a degree of polymerization (DP) from two to five. In this study, CtCDP was first discovered to have weak activities on weakly water-soluble celloheptaose and insoluble regenerated amorphous cellulose (RAC). To enhance its activity on solid cellulosic materials, four cellulose binding modules, e.g., CBM3 (type A) from C. thermocellum CbhA, CBM4-2 (type B) from Rhodothermus marinus Xyn10A, CBM6 (type B) from Cellvibrio mixtus Cel5B, and CBM9-2 (type C) from Thermotoga maritima Xyn10A, were fused to the C terminus of CtCDP. Fusion of any selected CBM with CtCDP did not influence its kinetic parameters on cellobiose but affected the binding and catalytic properties on celloheptaose and RAC differently. Among them, addition of CBM9 to CtCDP resulted in a 2.7-fold increase of catalytic efficiency for degrading celloheptaose. CtCDP-CBM9 exhibited enhanced specific activities over 20% on the short-chain RAC (DP = 14) and more than 50% on the long-chain RAC (DP = 164). The chimeric protein CtCDP-CBM9 would be the first step to construct a cellulose phosphorylase for in vitro hydrogen production from cellulose by synthetic pathway biotransformation (SyPaB).  相似文献   

6.
An amperometric enzyme biosensor for continuous detection of cellobiose has been implemented as an enzyme assay for cellulases. We show that the initial kinetics for cellobiohydrolase I, Cel7A from Trichoderma reesei, acting on different types of cellulose substrates, semi‐crystalline and amorphous, can be monitored directly and in real‐time by an enzyme‐modified electrode based on cellobiose dehydrogenase (CDH) from Phanerochaete chrysosporium (Pc). PcCDH was cross‐linked and immobilized on the surface of a carbon paste electrode which contained a mediator, benzoquinone. An oxidation current of the reduced mediator, hydroquinone, produced by the CDH‐catalyzed reaction with cellobiose, was recorded under constant‐potential amperometry at +0.5 V (vs. Ag/AgCl). The CDH‐biosensors showed high sensitivity (87.7 µA mM?1 cm?2), low detection limit (25 nM), and fast response time (t95% ~ 3 s) and this provided experimental access to the transient kinetics of cellobiohydrolases acting on insoluble cellulose. The response from the CDH‐biosensor during enzymatic hydrolysis was corrected for the specificity of PcCDH for the β‐anomer of cello‐oligosaccharides and the approach were validated against HPLC. It is suggested that quantitative, real‐time data on pure insoluble cellulose substrates will be useful in attempts to probe the molecular mechanism underlying enzymatic hydrolysis of cellulose. Biotechnol. Bioeng. 2012; 109: 3199–3204. © 2012 Wiley Periodicals, Inc.  相似文献   

7.
Oligosaccharides synthesized from a mixture of maltoheptaose and [U-13C]maltose with transglucosidase [EC 2.4.1.24] from Aspergillus niger were investigated. When the reaction mixture was incubated at 15 °C for 1 h, several types of oligosaccharides with DP (degree of polymerization) 2 to DP8 containing α-d-Glcp-(1→6)-maltoheptaose were detected by liquid chromatography-mass spectrometry (LC-MS) and methylation analysis. Most of these compounds consisted of α-(1→4) linkages in the main chain and α-(1→6) linkages at the non-reducing ends. However, when the reaction mixture was incubated for 96 h, most of these products were converted into oligosaccharides with DP2 to DP5 consisting of only α-(1→6) linkages. These results suggested that A. niger transglucosidase rapidly transferred glucosyl residues to maltooligosaccharides, and gradually hydrolyzed both α-(1→4) linkages and α-(1→6) linkages at the non-reducing end, and transformed these into smaller molecules of mainly α-(1→6) linkages.  相似文献   

8.
Inverting cellobiose phosphorylase (CtCBP) and cellodextrin phosphorylase (CtCDP) from Clostridium thermocellum ATCC27405 of glycoside hydrolase family 94 catalysed reverse phosphorolysis to produce cellobiose and cellodextrins in 57% and 48% yield from α-d-glucose 1-phosphate as donor with glucose and cellobiose as acceptor, respectively. Use of α-d-glucosyl 1-fluoride as donor increased product yields to 98% for CtCBP and 68% for CtCDP. CtCBP showed broad acceptor specificity forming β-glucosyl disaccharides with β-(1→4)- regioselectivity from five monosaccharides as well as branched β-glucosyl trisaccharides with β-(1→4)-regioselectivity from three (1→6)-linked disaccharides. CtCDP showed strict β-(1→4)-regioselectivity and catalysed linear chain extension of the three β-linked glucosyl disaccharides, cellobiose, sophorose, and laminaribiose, whereas 12 tested monosaccharides were not acceptors. Structure analysis by NMR and ESI-MS confirmed two β-glucosyl oligosaccharide product series to represent novel compounds, i.e. β-d-glucopyranosyl-[(1→4)-β-d-glucopyranosyl]n-(1→2)-d-glucopyranose, and β-d-glucopyranosyl-[(1→4)-β-d-glucopyranosyl]n-(1→3)-d-glucopyranose (n = 1–7). Multiple sequence alignment together with a modelled CtCBP structure, obtained using the crystal structure of Cellvibrio gilvus CBP in complex with glucose as a template, indicated differences in the subsite +1 region that elicit the distinct acceptor specificities of CtCBP and CtCDP. Thus Glu636 of CtCBP recognized the C1 hydroxyl of β-glucose at subsite +1, while in CtCDP the presence of Ala800 conferred more space, which allowed accommodation of C1 substituted disaccharide acceptors at the corresponding subsites +1 and +2. Furthermore, CtCBP has a short Glu496-Thr500 loop that permitted the C6 hydroxyl of glucose at subsite +1 to be exposed to solvent, whereas the corresponding longer loop Thr637–Lys648 in CtCDP blocks binding of C6-linked disaccharides as acceptors at subsite +1. High yields in chemoenzymatic synthesis, a novel regioselectivity, and novel oligosaccharides including products of CtCDP catalysed oligosaccharide oligomerisation using α-d-glucosyl 1-fluoride, all together contribute to the formation of an excellent basis for rational engineering of CBP and CDP to produce desired oligosaccharides.  相似文献   

9.
A fucoidan from the brown alga Fucus evanescens was effectively depolymerized by autohydrolysis. Negative-ion electrospray ionization mass spectrometry (ESIMS) revealed that the mixture contained sulfated mono- and oligosaccharides with polymerization degree (DP) up to 6, having from 1 to 4 sulfate groups per molecule. The prevalence of oligosaccharides with even DP was observed. It could be explained by the tendency of the 3-linked α-l-fucopyranose residues to hydrolyze faster than 4-linked ones. The intermolecular sulfate transfer during autohydrolysis was detected by ESIMS, when equimolar quantities of d-Rib and d-Glc were added as acceptors. The products were singly-sulfated and hexose was about four times more effective as an acceptor, than pentose. It was impossible to record MS/MS spectra of the sulfate transfer products, since intensities of their ions were too low.  相似文献   

10.
Cryptococcus sp. S-2 carboxymethyl cellulase (CSCMCase) is active in the acidic pH and lacks a binding domain. The absence of the binding domain makes the enzyme inefficient against insoluble cellulosic substrates. To enhance its binding affinity and its cellulolytic activity to insoluble cellulosic substrates, cellulose binding domain (CBD) of cellobiohydrolase I (CBHI) from Trichoderma reesei belonging to carbohydrate binding module (CBM) family 1 was fused at the C-terminus of CSCMCase. The constructed fusion enzymes (CSCMCase-CBD and CSCMCase-2CBD) were expressed in a newly recombinant expression system of Cryptococcus sp. S-2, purified to homogeneity, and then subject to detailed characterization. The recombinant fusion enzymes displayed optimal pH similar to those of the native enzyme. Compared with rCSCMCase, the recombinant fusion enzymes had acquired an increased binding affinity to insoluble cellulose and the cellulolytic activity toward insoluble cellulosic substrates (SIGMACELL® and Avicel) was higher than that of native enzyme, confirming the presence of CBDs improve the binding and the cellulolytic activity of CSCMCase on insoluble substrates. This attribute should make CSCMCase an attractive applicant for various application.  相似文献   

11.
When cells of Acanthamoeba castellanii are placed in a non-nutrient medium, they differentiate into cysts which possess cellulosic walls. In the present study, the source of the glucosyl unit for cyst wall cellulose was investigated by following the encystment of trophozoites grown in the presence of 14C-labeled fatty acids (uniformly labeled palmitate or oleate) or [3-3H]glucose. Cells were fractionated at the beginning and after 30 hr of encystment using a modified Schmidt-Tannhauser procedure. In cells grown on fatty acids, 90% of the labeled material was in the lipid fractions both before and after encystment with the total amount of label/cell changing very little. Both partial and complete acid hydrolysis of the glycogen of the acidsoluble fraction and the alkali-insoluble residue of the cyst wall indicated that the glucose of both fractions was not radioactive, although Acanthamoeba is known to have a functional glyoxylate pathway.Fractionation data of cells grown on [3H]glucose indicated a sevenfold increase in radioactivity in the wall insoluble fraction and a fivefold decrease in the acid-soluble fraction with the cpm/cell of the other fractions changing very little after 30 hr of encystment. Approximately 70% of the 3H-labeled material was recovered as glucose from the 30-hr wall insoluble fraction following complete acid hydrolysis. The specific radioactivity of glucose in the cyst wall insoluble fraction was the same as that of glycogen glucose isolated from the acid soluble fraction of trophozoites. Electron microscopic autoradiography showed that the majority of nonlipid radioactivity was due to glycogen in trophozoites. Autoradiograms failed to reveal Golgi bodies or any particular region of the cell as being the specialized site of cellulose synthesis. The results of the fractionation and autoradiographic studies are consistent with the concept that glycogen is a precursor of cyst wall cellulose, and that glucosyl units of glycogen and/or other glucose derivatives are converted to cellulose without significant dilution under the experimental conditions used.  相似文献   

12.
Soluble cellodextrins (linear β-1,4-d -gluco-oligosaccharides) have interesting applications as ingredients for human and animal nutrition. Their bottom-up synthesis from glucose is promising for bulk production, but to ensure a completely water-soluble product via degree of polymerization (DP) control (DP ≤ 6) is challenging. Here, we show biocatalytic production of cellodextrins with DP centered at 3 to 6 (~96 wt.% of total product) using coupled cellobiose and cellodextrin phosphorylase. The cascade reaction, wherein glucose was elongated sequentially from α-d -glucose 1-phosphate (αGlc1-P), required optimization and control at two main points. First, kinetic and thermodynamic restrictions upon αGlc1-P utilization (200 mM; 45°C, pH 7.0) were effectively overcome (53% → ≥90% conversion after 10 hrs of reaction) by in situ removal of the phosphate released via precipitation with Mg2+. Second, the product DP was controlled by the molar ratio of glucose/αGlc1-P (∼0.25; 50 mM glucose) used in the reaction. In optimized conversion, soluble cellodextrins in a total product concentration of 36 g/L were obtained through efficient utilization of the substrates used (glucose: 98%; αGlc1-P: ∼80%) after 1 hr of reaction. We also showed that, by keeping the glucose concentration low (i.e., 1–10 mM; 200 mM αGlc1-P), the reaction was shifted completely towards insoluble product formation (DP ∼9–10). In summary, this study provides the basis for an efficient and product DP-controlled biocatalytic synthesis of cellodextrins from expedient substrates.  相似文献   

13.
Xyloglucan oligosaccharides were isolated with various degreesof polymerization (DP) and reduced with tritiated sodium borohydride.The 3H-oligosaccharides were tested for their ability to bindto amorphous and microcrystalline celluloses and to cellulosefilter paper. The time course of binding indicated that theradiolabeled oligosaccharides continued to be bound for at least1 h after heating at 120°C. The binding probably requiredthe organization of the oligosaccharides and celluloses by gradualannealing after heating. Although neither pentasaccharide (glucose:xylose, 3 : 2), heptasaccharide (glucose: xylose, 4 : 3) andnonasaccharide (glucose : xylose : galactose : fucose, 4 : 3: 1 : 1) failed to bind to the celluloses, binding occurredwith oligosaccharides with DP equivalent to more than four consecutive1,4-ß-glucosyl residues. The extent of binding tothe celluloses increased gradually from octasaccharide (glucose:xylose, 5 : 3) to hendecosanosaccharide (glucose/xylose, 12: 9), with the increase in the DP of 1,4-ß-glucosylresidues. The binding of reduced cello-dextrins to celluloserequired at least 4 consecutive 1,4-ß-glucosyl residues.The extent of binding of cellopentitol or cellohexitol to cellulosewas similar to that of hendecosanosaccharide, showing lowerbinding for xyloglucan oligosaccharides in spite of longer chainsof 1,4-ß-glucosyl residues. These findings suggestthat the mode of binding to cellulose of xyloglucan oligosaccharidesis different from that of cello-oligosaccharides. (Received February 18, 1994; Accepted June 1, 1994)  相似文献   

14.
During growth in liquid culture medium, that contained single soluble or insoluble cellulosic carbon source, Aspergillus sydowi (Bain. & Start.) Thom & Church released cellulolytic enzymes into the medium. The enzymes were separated by gel filtration followed by ion exchange chromatography into three components, all of high molecular weight. One of the components (Ac) has the character of a C1 cellulase enzyme. In the assay for hydrolysis of insoluble cellulose, the combined fractions, especially whenever the fraction under test contained the component Ac, released more glucose than when each component was employed alone.  相似文献   

15.
Glucooligosaccharides (GOS) were produced by using an acceptor reaction with two types of glucansucrase (GTF-S and GTF-I) from Streptococcus sobrinus. Acceptor reactions of GTF-S with maltose acceptor, gave a great number of GOS ranging from DP(degree of polymerization) 2 to DP15. At the both acceptor reactions with GTF-S or GTF-I, as the sucrose/maltose ratio was decreased, the amount of dextran and DP of oligosaccharides was decreased. A maximum GOS yield of 69% was achieved at the acceptor reaction with GTF-I and when the molar ratio of sucrose/maltose is 2:1, in which GOS of DP6~DP9 were major oligosaccharides and 17% of dextran. The polymeric size of GOS could be controlled by varying the ratio of sucrose to the acceptor (maltose in this work).  相似文献   

16.
Extracellular dextranases were extracted from a dextran-degrading microorganism, Bacteroides oralis Ig4a, which had been isolated from human dental plaque, and purified. Crude enzyme preparations obtained from a broth culture supernatant by salting out with ammonium sulfate were subjected to column chromatography on DEAE-cellulose and subsequent Bio-Gel p-100, followed by isoelectric focusing. Two kinds of enzyme preparations, Enzymes I and II, with the ability to degrade soluble dextran were obtained. The optimal pHs of Enzymes I and II were 5.5 and 6.8, and the isoelectric points were pH 4.5 and 6.5, respectively. The molecular weights of Enzymes I and II were estimated by SDS-PAGE to be 44,000 and 52,000. Both enzymes were inhibited by Pb2+ and Fe3+, but not by Ca2+, Mg2+, Zn2+, or Fe2+. Neither the presence of EDTA nor iodoacetamide had any appreciable effect on the enzyme activity. The enzyme activity was independent of any of these metal ions. Enzyme I liberated glucose, isomaltose, maltotriose and higher oligosaccharides from dextran. In contrast, Enzyme II liberated only glucose from dextran and was assumed to be an exoglycosidase. Neither of the enzymes degraded modified insoluble glucan, which is a partially oxidized mutan of S. mutans containing predominantly α-(1, 3) linkages.  相似文献   

17.
Glycogen, highly branched (1→4)(1→6)-linked α-d-glucan, can be extracted from natural sources such as animal tissues or shellfish (natural source glycogen, NSG). Glycogen can also be synthesized in vitro from glucose-1-phosphate using the cooperative action of α-glucan phosphorylase (GP, EC 2.4.1.1) and branching enzyme (BE, EC 2.4.1.18), or from short-chain amylose by the cooperative action of BE and amylomaltase (AM, EC 2.4.1.25). It has been shown that enzymatically synthesized glycogen (ESG) has structural and physicochemical properties similar to those of NSG. In this study, the fine structures of ESG and NSG were analyzed using isoamylase and α-amylase. Isoamylase completely hydrolyzed the α-1,6 linkages of ESG and NSG. The unit-chain distribution (distribution of degrees of polymerization (DP) of α-1,4 linked chains) of ESG was slightly narrower than that of NSG. α-Amylase treatment revealed that initial profiles of hydrolyses of ESG and NSG were almost the same: both glycogens were digested slowly, compared with starch. The final products from NSG by α-amylase hydrolysis were glucose, maltose, maltotriose, branched oligosaccharides with DP ? 4, and highly branched macrodextrin molecules with molecular weights of up to 10,000. When ESG was digested with excess amounts of α-amylase, much larger macrodextrins (molecular weight > 106) were detected. In contrast, oligosaccharides with DP 4-7 could not be detected from ESG. These results suggest that the α-1,6 linkages in ESG molecules are more regularly distributed than those in NSG molecules.  相似文献   

18.
The transglycosylation activity of barley α-amylase 1 (AMY1) and active site AMY1 subsite mutant enzymes was investigated. We report here the transferase ability of the V47A, V47F, V47D and S48Y single mutants and V47K/S48G and V47G/S48D double mutant AMY1 enzymes in which the replaced amino acids play important role in substrate binding at subsites at −3 through −5. Although mutation increases the transglycosylation activity of enzymes, in the presence of acceptors the difference between wild type and mutants is not so significant. Oligomer transfer reactions of AMY1 wild type and its mutants were studied using maltoheptaose and maltopentaose donors and different chromophore containing acceptors. The conditions for the chemoenzymatic synthesis of 4-methylumbelliferyl-α-d-maltooligosaccharides (MU-α-d-MOSs) were optimized using 4-methylumbelliferyl-β-d-glucoside as acceptor and maltoheptaose as donor. 4-Methylumbelliferyl-α-d-maltoside, -maltotrioside, -maltotetraoside and -maltopentaoside have been synthesized. Products were identified by MALDI-TOF MS. 1H and 13C NMR analyses showed that AMY1 V47F preserved the stereo- and regioselectivity. The produced MU-α-d-MOSs of degree of polymerization DP 2, DP 3 and DP 5 were successfully applied to detect activity of Bacillus stearothermophilus maltogenic α-amylase, human salivary α-amylase and Bacillus licheniformis α-amylase, respectively in a fast and simple fluorometric assay.  相似文献   

19.
Transglycosylation reactions are useful for preserving a specific sugar structure during the synthesis of branched oligosaccharides. We have previously reported a panosyl unit transglycosylation reaction by pullulan-hydrolyzing amylase II (TVA II) cloned from Thermoactinomyces vulgaris R-47 (Tonozuka et al., Carbohydr. Res., 1994, 261, 157–162). The acceptor specificity of the TVA II transglycosylation reaction was investigated using pullulan as the donor and sugar alcohols as the acceptor. TVA II transferred the α-panosyl unit to the C-1 hydroxyl group of meso-erythritol, C-1 and C-2 of xylitol, and C-1 and C-6 of d-sorbitol. TVA II differentiated between the sugar alcohols’ hydroxyl groups to produce five novel non-reducing branched oligosaccharides, 1-O-α-panosylerythritol, 1-O-α-panosylxylitol, 2-O-α-panosylxylitol, 1-O-α-panosylsorbitol, and 6-O-α-panosylsorbitol. The Trp356→Ala mutant showed similar transglycosylation reactions; however, panose production by the mutant was 4.0–4.5-fold higher than that of the wild type. This suggests that Trp356 is important for recognizing both water and the acceptor molecules in the transglycosylation and the hydrolysis reaction.  相似文献   

20.
Glyphosate with an equivalent concentration of either 0, 2.16 or 8.64 kg/hm2 was sprayed on to cellulosic materials before burying in two soil types; peat (soil I) and sandy clay loam (soil II). Alternatively the soils were sprayed with 0, 20 or 150 ppm of the herbicide before burying the cellulosic material either immediately or after preincubation for 4 weeks. In soil I, the increase in glyphosate concentrations substantially reduced the decomposition of cellulosic material regardless of the method of application employed. Glyphosate at 8.64 kg/hm2 reduced the mass loss of the treated substrate by 83%. However, cellulose decomposition in soil preincubated for 4 weeks before burying was affected almost to the same extent as the untreated control. Glyphosate stimulated cellulose decomposition when substrates were buried in soil II. Mass loss in soil treated with 150 ppm increased by about 100% while when glyphosate was sprayed directly to the substrate (at 8.64 kg/hm2), the loss was about 25%.  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