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 21 毫秒
1.
Two Streptomyces strains, S. viridosporus T7A and S. setonii 75Vi2, were grown on softwood, hardwood, and grass lignocelluloses, and lignocellulose decomposition was followed by monitoring substrate weight loss, lignin loss, and carbohydrate loss over time. Results showed that both Streptomyces strains substantially degraded both the lignin and the carbohydrate components of each lignocellulose; however, these actinomycetes were more efficient decomposers of grass lignocelluloses than of hardwood or softwood lignocelluloses. In particular, these Streptomyces strains were more efficient decomposers of grass lignins than of hardwood or softwood lignins.  相似文献   

2.
Summary Serratia marcescens was found to degrade kraft lignin by only 15%. When 14C-radiolabelled lignocelluloses and DHP lignins were used as substrates the bacterium mineralized to 14CO2 only 1.1–1.9% and 0.4–0.8% of the lignins respectively. However, some 44.4% of the 14C--DHP lignin was recovered as soluble radiolabelled products.  相似文献   

3.

Background

For cellulosic biofuels processes, suitable characterization of the lignin remaining within the cell wall and correlation of quantified properties of lignin to cell wall polysaccharide enzymatic deconstruction is underrepresented in the literature. This is particularly true for grasses which represent a number of promising bioenergy feedstocks where quantification of grass lignins is particularly problematic due to the high fraction of p-hydroxycinnamates. The main focus of this work is to use grasses with a diverse range of lignin properties, and applying multiple lignin characterization platforms, attempt to correlate the differences in these lignin properties to the susceptibility to alkaline hydrogen peroxide (AHP) pretreatment and subsequent enzymatic deconstruction.

Results

We were able to determine that the enzymatic hydrolysis of cellulose to to glucose (i.e. digestibility) of four grasses with relatively diverse lignin phenotypes could be correlated to total lignin content and the content of p-hydroxycinnamates, while S/G ratios did not appear to contribute to the enzymatic digestibility or delignification. The lignins of the brown midrib corn stovers tested were significantly more condensed than a typical commercial corn stover and a significant finding was that pretreatment with alkaline hydrogen peroxide increases the fraction of lignins involved in condensed linkages from 88?C95% to ~99% for all the corn stovers tested, which is much more than has been reported in the literature for other pretreatments. This indicates significant scission of ??-O-4 bonds by pretreatment and/or induction of lignin condensation reactions. The S/G ratios in grasses determined by analytical pyrolysis are significantly lower than values obtained using either thioacidolysis or 2DHSQC NMR due to presumed interference by ferulates.

Conclusions

It was found that grass cell wall polysaccharide hydrolysis by cellulolytic enzymes for grasses exhibiting a diversity of lignin structures and compositions could be linked to quantifiable changes in the composition of the cell wall and properties of the lignin including apparent content of the p-hydroxycinnamates while the limitations of S/G estimation in grasses is highlighted.  相似文献   

4.

Background

Lignin is embedded in the plant cell wall matrix, and impedes the enzymatic saccharification of lignocellulosic feedstocks. To investigate whether enzymatic digestibility of cell wall materials can be improved by altering the relative abundance of the two major lignin monomers, guaiacyl (G) and syringyl (S) subunits, we compared the degradability of cell wall material from wild-type Arabidopsis thaliana with a mutant line and a genetically modified line, the lignins of which are enriched in G and S subunits, respectively.

Results

Arabidopsis tissue containing G- and S-rich lignins had the same saccharification performance as the wild type when subjected to enzyme hydrolysis without pretreatment. After a 24-hour incubation period, less than 30% of the total glucan was hydrolyzed. By contrast, when liquid hot water (LHW) pretreatment was included before enzyme hydrolysis, the S-lignin-rich tissue gave a much higher glucose yield than either the wild-type or G-lignin-rich tissue. Applying a hot-water washing step after the pretreatment did not lead to a further increase in final glucose yield, but the initial hydrolytic rate was doubled.

Conclusions

Our analyses using the model plant A. thaliana revealed that lignin composition affects the enzymatic digestibility of LHW pretreated plant material. Pretreatment is more effective in enhancing the saccharification of A. thaliana cell walls that contain S-rich lignin. Increasing lignin S monomer content through genetic engineering may be a promising approach to increase the efficiency and reduce the cost of biomass to biofuel conversion.  相似文献   

5.
Decomposition of lignocelluloses from Spartina alterniflora in salt-marsh sediments was measured by using 14C-labeled compounds. Rates of decomposition were fastest in the first 4 days of incubation and declined later. Lignins labeled in side chains were mineralized slightly faster than uniformly labeled lignins; 12% of the [side chain-14C]lignin-labeled lignocellulose was mineralized after 816 h of incubation, whereas only 8% of the [U-14C]lignin-labeled lignocelluloses were degraded during this period. The carbohydrate moiety within the lignocellulose complex was degraded about four times faster than the lignin moiety; after 816 h of incubation, 29 to 37% of the carbohydrate moiety had been mineralized. Changes in concentration of lignin and cellulose in litter of S. alterniflora were followed over 2 years of decay. Cellulose disappeared from litter more rapidly than lignin; 50% of the initial content of cellulose was lost after 130 days, whereas lignin required 330 to 380 days for 50% loss. The slow loss of lignin compared with other litter components resulted in a progressive enrichment of litter in lignin content. The rates of mineralization of [14C]lignocelluloses in marsh sediments were similar to the rates of lignocellulose decomposition in litter on the marsh.  相似文献   

6.
Molecular markers were used to characterise sugarcane intergeneric hybrids between S. officinarum and E. arundinaceus. Very simple diagnostic tools for hybrid identification among the progeny were derived from isozyme electrophoresis and a sequence-tagged PCR. Two enzyme systems (GOT and MDH B) and PCR amplification revealing spacer-size variation in the 5s-rDNA cluster were found most convenient. Specific characterisation of the two genomic components was possible using RFLP and in situ hybridisation. The strong molecular differentiation between S. officinarum and E. arundinaceus allows the identification of numerous Erianthus-specific RFLP bands in the hybrids. Genomic DNA in situ hybridisation allows for the differentiation of the chromosomes contributed by S. officinarum and E. arundinaceus in chromosome preparations of the hybrids. In situ hybridisation with the 18s-5.8s-25s rDNA probe highlights the basic chromosome numbers in the two parental species. The potential of these techniques to monitor the Erianthus genome during the introgression process is discussed.  相似文献   

7.
The aromatic composition of lignin is an important trait that greatly affects the usability of lignocellulosic biomass. We previously identified a rice (Oryza sativa) gene encoding coniferaldehyde 5‐hydroxylase (OsCAld5H1), which was effective in modulating syringyl (S)/guaiacyl (G) lignin composition ratio in rice, a model grass species. Previously characterized OsCAld5H1‐knockdown rice lines, which were produced via an RNA‐interference approach, showed augmented G lignin units yet contained considerable amounts of residual S lignin units. In this study, to further investigate the effect of suppression of OsCAld5H1 on rice lignin structure, we generated loss‐of‐function mutants of OsCAld5H1 using the CRISPR/Cas9‐mediated genome editing system. Homozygous OsCAld5H1‐knockout lines harboring anticipated frame‐shift mutations in OsCAld5H1 were successfully obtained. A series of wet‐chemical and two‐dimensional NMR analyses on cell walls demonstrated that although lignins in the mutant were predictably enriched in G units all the tested mutant lines produced considerable numbers of S units. Intriguingly, lignin γ‐p‐coumaroylation analysis by the derivatization followed by reductive cleavage method revealed that enrichment of G units in lignins of the mutants was limited to the non‐γ‐p‐coumaroylated units, whereas grass‐specific γ‐p‐coumaroylated lignin units were almost unaffected. Gene expression analysis indicated that no homologous genes of OsCAld5H1 were overexpressed in the mutants. These data suggested that CAld5H is mainly involved in the production of non‐γ‐p‐coumaroylated S lignin units, common in both eudicots and grasses, but not in the production of grass‐specific γ‐p‐coumaroylated S units in rice.  相似文献   

8.
In the context of our research on cell wall formation and maturation in flax (Linum usitatissimum L) bast fibers, we (1) confirmed the presence of lignin in bast fibers and (2) quantified and characterized the chemical nature of this lignin at two developmental stages. Histochemical methods (Weisner and Maüle reagents and KMnO4-staining) indicating the presence of lignin in bast fibers at the light and electron microscope levels were confirmed by chemical analyses (acetyl bromide). In general, the lignin content in flax bast fibers varied between 1.5% and 4.2% of the dry cell wall residues (CWRs) as compared to values varying between 23.7% and 31.4% in flax xylem tissues. Immunological and chemical analyses (thioacidolysis and nitrobenzene oxidation) indicated that both flax xylem- and bast fiber-lignins were rich in guaiacyl (G) units with S/G values inferior to 0.5. In bast fibers, the highly sensitive immunological probes allowed the detection of condensed guaiacyl-type (G) lignins in the middle lamella, cell wall junctions, and in the S1 layer of the secondary wall. In addition, lower quantities of mixed guaiacyl–syringyl (GS) lignins could be detected throughout the secondary cell wall. Chemical analyses suggested that flax bast-fiber lignin is more condensed than the corresponding xylem lignin. In addition, H units represented up to 25% of the monomers released from bast-fiber lignin as opposed to a value of 1% for the corresponding xylem tissue. Such an observation indicates that the structure of flax bast-fiber lignin is significantly different from that of the more typical woody plant lignin, thereby suggesting that flax bast fibers represent an interesting system for studying an unusual lignification process.  相似文献   

9.
Pan YB  Burner DM  Legendre BL 《Genetica》2000,108(3):285-295
5S rRNA intergenic spacers were amplified from two elite sugarcane (Saccharumhybrids) cultivars and their related taxa by polymerase chain reaction (PCR) with 5S rDNA consensus primers. Resulting PCR products were uniform in length from each accession but exhibited some degree of length variation among the sugarcane accessions and related taxa. These PCR products did not always cross hybridize in Southern blot hybridization experiments. These PCR products were cloned into a commercial plasmid vector PCR™ 2.1 and sequenced. Direct sequencing of cloned PCR products revealed spacer length of 231–237 bp for S. officinarum, 233–237 for sugarcane cultivars, 228–238 bp for S. spontaneum, 239–252 bp for S. giganteum, 385–410 bp for Erianthusspp., 226–230 bp for Miscanthus sinensisZebra, 206–207 bp for M. sinensisIMP 3057, 207–209 bp for Sorghum bicolor, and 247–249 bp for Zea mays. Nucleotide sequence polymorphism were found at both the segment and single nucleotide level. A consensus sequence for each taxon was obtained by Align X. Multiple sequences were aligned and phylogenetic trees constructed using Align X, CLUSTAL and DNAMAN programs. In general, accessions of the following taxa tended to group together to form distinct clusters: S. giganteum, Erianthusspp., M. sinensis, S. bicolor, and Z. mays. However, the two S. officinarumclones and two sugarcane cultivars did not form distinct clusters but interrelated within the S. spontaneumcluster. The disclosure of these 5S rRNA intergenic spacer sequences will facilitate marker-assisted breeding in sugarcane. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

10.
Specifically radiolabeled [14C-lignin]lignocelluloses were prepared from the aquatic macrophytes Spartina alterniflora, Juncus roemerianus, Rhizophora mangle, and Carex walteriana by using [14C]phenylalanine, [14C]tyrosine, and [14C]cinnamic acid as precursors. Specifically radiolabeled [14C-polysaccharide]lignocelluloses were prepared by using [14C]glucose as precursor. The rates of microbial degradation varied among [14C-lignin]lignocelluloses labeled with different lignin precursors within the same plant species. To determine the causes of these differential rates, [14C-lignin]lignocelluloses were thoroughly characterized for the distribution of radioactivity in nonlignin contaminants and within the lignin macromolecule. In herbaceous plants, significant amounts (8 to 24%) of radioactivity from [14C]phenylalanine and [14C]tyrosine were found associated with protein, although very little (3%) radioactivity from [14C]cinnamic acid was associated with protein. Microbial degradation of radiolabeled protein resulted in overestimation of lignin degradation rates in lignocelluloses derived from herbaceous aquatic plants. Other differences in degradation rates among [14C-lignin]lignocelluloses from the same plant species were attributable to differences in the amount of label being associated with ester-linked subunits of peripheral lignin. After acid hydrolysis of [14C-polysaccharide]lignocelluloses, radioactivity was detected in several sugars, although most of the radioactivity was distributed between glucose and xylose. After 576 h of incubation with salt marsh sediments, 38% of the polysaccharide component and between 6 and 16% of the lignin component (depending on the precursor) of J. roemerianus lignocellulose was mineralized to 14CO2; during the same incubation period, 30% of the polysaccharide component and between 12 and 18% of the lignin component of S. alterniflora lignocellulose was mineralized.  相似文献   

11.
Two Streptomyces strains were grown on sugarcane bagasse and groundnut hulls lignocelluloses in semi-solid state culture at 37°C for 12 weeks. Best results gave a 45% depletion of sugarcane bagasse lignocellulose with a 21% crude protein content of final material. The possibility of using S. viridosporus to improve the protein content of both lignocelluloses for use as an animal feedstock supplement is discussed.At the time of this research the authors were with the Department of Biological Sciences, Rivers State University of Science & Technology, PMB 5080, Port Harcourt, Nigeria and the Department of Biological Sciences, University of Calabar, PMB 1115, Calabar, Nigeria. Dr lyo is now with the Department of Microbiology, University of Guelph, Guelph, Ontario N1G 2W1, Canada.  相似文献   

12.
Decomposition of lignocelluloses from Spartina alterniflora in salt-marsh sediments was measured by using C-labeled compounds. Rates of decomposition were fastest in the first 4 days of incubation and declined later. Lignins labeled in side chains were mineralized slightly faster than uniformly labeled lignins; 12% of the [side chain-C]lignin-labeled lignocellulose was mineralized after 816 h of incubation, whereas only 8% of the [U-C]lignin-labeled lignocelluloses were degraded during this period. The carbohydrate moiety within the lignocellulose complex was degraded about four times faster than the lignin moiety; after 816 h of incubation, 29 to 37% of the carbohydrate moiety had been mineralized. Changes in concentration of lignin and cellulose in litter of S. alterniflora were followed over 2 years of decay. Cellulose disappeared from litter more rapidly than lignin; 50% of the initial content of cellulose was lost after 130 days, whereas lignin required 330 to 380 days for 50% loss. The slow loss of lignin compared with other litter components resulted in a progressive enrichment of litter in lignin content. The rates of mineralization of [C]lignocelluloses in marsh sediments were similar to the rates of lignocellulose decomposition in litter on the marsh.  相似文献   

13.
Grass lignocelluloses, such as those in corn and switchgrass, are a major resource in the emerging cellulose-to-ethanol strategy for biofuels. The potential bioconversion of carbohydrates in this potential resource, however, is limited by the associated aromatic constituents within the grass fiber. These aromatics include both lignins, which are phenylpropanoid units of various types, and low-molecular weight phenolic acids. Structural and chemical studies over the years have identified the location and limitation to fiber degradation imposed by a variety of these aromatic barriers. For example, coniferyl lignin appears to be the most effective limitation to biodegradation, existing in xylem cells of vascular tissues. On the other hand, cell walls with syringyl lignin, e.g., leaf sclerenchyma, are often less recalcitrant. Ferulic and p-coumaric acids that are esterified to hemicellulosic sugars constitute a major limitation to biodegradation in non-lignified cell walls in grass fibers, especially warm season species. Non-chemical methods to improve bioconversion of the lignocelluloses through modification of aromatics include: (1) use of lignin-degrading white rot fungi, (2) pretreatment with phenolic acid esterases, and (3) plant breeding to modify cell wall aromatics. In addition to increased availability of carbohydrates for fermentation, separation and collection of aromatics could provide value-added co-products to improve the economics of bioconversion. JIMB-2008: BioEnergy—Special issue.  相似文献   

14.
The ability of Streptomyces ipomoea laccase to polymerize secoisolariciresinol lignan and technical lignins was assessed. The reactivity of S. ipomoea laccase was also compared to that of low redox fungal laccase from Melanocarpus albomyces using low molecular mass p-coumaric, ferulic and sinapic acid as well as natural (acetosyringone) and synthetic 2,2,6,6-tetramethylpiperidine 1-oxyl (TEMPO) mediators as substrates. Oxygen consumption measurement, MALDI-TOF MS and SEC were used to follow the enzymatic reactions at pH 7, 8, 9 and 10 at 30 °C and 50 °C. Polymerization of lignins and lignan by S. ipomoea laccase under alkaline reaction conditions was observed, and was enhanced in the presence of acetosyringone almost to the level obtained with M. albomyces laccase without mediator. Reactivities of the enzymes towards acetosyringone and TEMPO were similar, suggesting exploitation of the compounds and low redox laccase in lignin valorization under alkaline conditions. The results have scientific impact on basic research of laccases.  相似文献   

15.
Specifically radiolabeled [14C-lignin]lignocelluloses and [14C-polysaccharide]lignocelluloses were prepared from a variety of marine and freshwater wetland plants including a grass, a sedge, a rush, and a hardwood. These [14C]lignocellulose preparations and synthetic [14C]lignin were incubated anaerobically with anoxic sediments collected from a salt marsh, a freshwater marsh, and a mangrove swamp. During long-term incubations lasting up to 300 days, the lignin and polysaccharide components of the lignocelluloses were slowly degraded anaerobically to 14CO2 and 14CH4. Lignocelluloses derived from herbaceous plants were degraded more rapidly than lignocellulose derived from the hardwood. After 294 days, 16.9% of the lignin component and 30.0% of the polysaccharide component of lignocellulose derived from the grass used (Spartina alterniflora) were degraded to gaseous end products. In contrast, after 246 days, only 1.5% of the lignin component and 4.1% of the polysaccharide component of lignocellulose derived from the hardwood used (Rhizophora mangle) were degraded to gaseous end products. Synthetic [14C]lignin was degraded anaerobically faster than the lignin component of the hardwood lignocellulose; after 276 days, 3.7% of the synthetic lignin was degraded to gaseous end products. Contrary to previous reports, these results demonstrate that lignin and lignified plant tissues are biodegradable in the absence of oxygen. Although lignocelluloses are recalcitrant to anaerobic biodegradation, rates of degradation measured in aquatic sediments are significant and have important implications for the biospheric cycling of carbon from these abundant biopolymers.  相似文献   

16.
Summary The wood-decay fungi Coriolus versicolor, a white-rot fungus, and Poria placenta, a brown-rot fungus, were grown on an extractive-free lignocellulose prepared from quackgrass (Agropyron repens). Their abilities to decompose this lignocellulose were compared to their abilities to decompose softwood (Picea pungens) and hardwood (Acer rubrum) lignocelluloses. The two fungi were grown on malt-extract dampened lignocelluloses at 28°C for up to 12 weeks. Replicate cultures were periodically harvested and lignocellulose decomposition was followed by monitoring substrate weight loss, lignin loss, and carbohydrate loss. Coriolus versicolor decomposed the lignin and carbohydrate components of the grass lignocellulose as efficiently as the softwood and hardwood lignocelluloses. Poria placenta, however, was not an efficient degrader of either lignin or carbohydrate in the grass lignocellulose. Poria placenta readily decomposed carbohydrate components of the softwood lignocellulose but not the hardwood lignocellulose.Paper number 81520 of the Idaho Agricultural Experiment Station  相似文献   

17.
Variation at the 18S+26S and 5S ribosomal DNA loci was assessed on 62 Erianthus Michx. clones, representing 11 species, and 15 clones from two Saccharum L. species used as a reference. Genus-specific markers for Erianthus Michx. sect. Ripidium Henrard (Old World species) were identified. Ribosomal DNA units in Erianthus sect. Ripidium exhibited an additional BamHI site compared to Saccharum, and 5S units showed length and restriction-site differences between Erianthus and Saccharum. These markers will be useful to follow introgression in Saccharum x Erianthus hybrids. Six ribosomal units (for 18+26S genes) were revealed in Erianthus sect. Ripidium, differing by restriction-site positions and/or length. These results provided new information on species relationships and evolution within the genus Erianthus. The Indonesian and Indian forms of E. arundinaceus (Retz.) Jeswiet gave different restriction patterns, which were similar to those of E. bengalense (Retz.) R. C. Bharadwaja and E. procerus (Roxb.) Raizade, respectively. The two 2n=20 species, E. ele-phantinus Hook.f. and E. ravennae (L.) P. Beauv., could also be differentiated at this locus. Two of the New World Erianthus species studied, E. rufipilus (Steud.) Griseb. and E. longisetosus Andersson, appeared more like Erianthus sect. Ripidium, whereas E. trinii Hack, and E. brevibardis Michx. showed patterns consistent with Miscanthus sinensis Andersson and S. spontaneum L., respectively. Finally, the comparison of rDNA restriction maps among Erianthus sect. Ripidium, Saccharum, sorghum and maize, led to unexpected conclusions concerning the relationships between the different genera and the position of Erianthus in the Saccharum complex.  相似文献   

18.
A cDNA clone encoding the lignin‐related enzyme caffeoyl CoA 3‐O‐methyltransferase (CCoAOMT) was isolated from a Pinus radiata cDNA library derived from differentiating xylem. Suppression of PrCCoAOMT expression in P. radiata tracheary element cultures affected lignin content and composition, resulting in a lignin polymer containing p‐hydroxyphenyl (H), catechyl (C) and guaiacyl (G) units. Acetyl bromide‐soluble lignin assays revealed reductions in lignin content of up to 20% in PrCCoAOMT‐deficient transgenic lines. Pyrolysis‐GC/MS and 2D‐NMR studies demonstrated that these reductions were due to depletion of G‐type lignin. Correspondingly, the proportion of H‐type lignin in PrCCoAOMT‐deficient transgenic lines increased, resulting in up to a 10‐fold increase in the H/G ratio relative to untransformed controls. 2D‐NMR spectra revealed that PrCCoAOMT suppression resulted in formation of benzodioxanes in the lignin polymer. This suggested that phenylpropanoids with an ortho‐diphenyl structure such as caffeyl alcohol are involved in lignin polymerization. To test this hypothesis, synthetic lignins containing methyl caffeate or caffeyl alcohol were generated and analyzed by 2D‐NMR. Comparison of the 2D‐NMR spectra from PrCCoAOMT‐RNAi lines and synthetic lignins identified caffeyl alcohol as the new lignin constituent in PrCCoAOMT‐deficient lines. The incorporation of caffeyl alcohol into lignin created a polymer containing catechyl units, a lignin type that has not been previously identified in recombinant lignin studies. This finding is consistent with the theory that lignin polymerization is based on a radical coupling process that is determined solely by chemical processes.  相似文献   

19.
Lignins result from the oxidative polymerization of three hydroxycinnamyl (p‐coumaryl, coniferyl and sinapyl) alcohols in a reaction mediated by peroxidases (EC 1.11.1.7) and laccases (EC 1.10.3.2), yielding H, G and S units, respectively. Although both acidic and basic peroxidases can oxidize p‐coumaryl and coniferyl alcohol, only basic peroxidases are able to oxidize sinapyl alcohol. The AtPrx52 from Arabidopsis is a basic peroxidase that has been reported to be highly homologous to the basic peroxidase of Zinnia elegans, the only peroxidase which has been unequivocally linked to lignin formation. Here, we show how the suppression of AtPrx52 causes a change in lignin composition, mainly at the level of stem interfascicular fibers. Quantification of lignins in two different atprx52 knock‐out mutants revealed a decrease of lignin amount compared with wild type. The S/G ratio, obtained by both nitrobenzene oxidation and thioacidolysis, indicated a decrease in S units in the atprx52 mutants. As deduced from Wiesner and mainly Mäule staining, this reduction in S unit content appears to be restricted to the interfascicular fibers. Moreover, quantitative polymerase chain reaction analysis in atprx52 plants showed a general downregulation of genes involved in lignin biosynthetic pathway, as well as genes related to secondary cell wall. On the other hand, other routes from phenylpropanoid metabolism were induced. Taken together, our results indicate that AtPrx52 is involved in the synthesis of S units in interfascicular fibers at late stages of the lignification process.  相似文献   

20.
Lignin, one of the main structural polymer of plant cell walls, varies in amount and monomeric composition among tissue and cell types, as well as among plant species. However, few analytical methods are available that can conveniently and accurately determine the morphological distribution of lignin units at the cellular level. In this report, we used time-of-flight secondary ion mass spectrometry (TOF-SIMS) to directly map guaiacyl (G) and syringyl (S) lignin units in several successive growth rings of the maple xylem. TOF-SIMS imaging and a semiquantitative approach revealed clear difference in the annual distribution of lignins between the fiber and vessel. While the vessel walls were constantly G-rich with varied S/G ratios through a growth ring, the fibers showed fairly regular annual distribution of lignins in which the earlywood was S-rich with an almost constant S/G ratio and the latewood was G-rich resulting from a decrease of the S unit. The reliability of TOF-SIMS results was demonstrated by its high correlation with the results of thioacidolysis on radial distribution of the S/G ratio in several contiguous tree rings and also in the latewood and earlywood of each ring. These results indicate that TOF-SIMS allows direct visualization of lignin composition in plant tissues.  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