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.
The monomeric composition of the suberins from 16 species of higher plants was determined by chromatographic methods following depolymerization of the isolated extractive-free cork layers with sodium methoxide-methanol. 1-Alkanols (mainly C18C28), alkanoic (mainly C16C30), α,ω-alkanedioic (mainly C16C24), ω-hydroxyalkanoic (mainly C16C21), dihydroxyhexadecanoic (mainly 10,16-dihydroxy- and 16-dihydroxyhexadecanoic), monohydroxyepoxyalkanoic (9,10-epoxy-18-hydroxyoctadecanoic), trihydroxyalkanoic (9,10, 18-trihydroxyoctadecanoic), epoxyalkanedioic (9,10-epoxyoctadecane-1,18-dioic) and dihydroxyalkanedioic (9,10-dihydroxyoctadecane-1 18-dioic) acids were detected in all species. The suberins differed from one another mainly in the relative proportions of these monomer classes and in the homologue content of their 1-alkanol, alkanoic, α,ω-alkanedioic and ω-hydroxyalkanoic acid fractions. C18 epoxy and vic-diol monomers were major components (32–59%) of half of the suberins examined (Quercus robur, Q. ilex, Q. suber, Fagus sylvatica, Castanea sativa, Betula pendula, Acer griseum, Fraxinus excelsior) where as ω-hydroxyalkanoic and α,ω-alkanedioic acids predominated in those that contained smaller quantities of such polar C18 monomers (Acer pseudoplatanus, Ribes nigrum, Euonymus alatus, Populus tremula, Solanum tuberosum, Sambucus nigra, Laburnum anagyroides, Cupressus leylandii). All species, however, contained substantial amounts (14–55 %) of ω-hydroxyalkanoic acids, the most common homologues being 18:1 (9) and 22: 0. The dominant α,ω-alkanedioic acid homologues were 16: 0 and 18: 1 (9) whereas 22: 0, 24: 0 and 26: 0, and 20: 0, 22: 0 and 24: 0 were usually the principal homologues in the 1-alkanol and alkanoic acid fractions, respectively. The most diagnostic feature of the suberins examined was the presence of monomers greater than C18 in chain length; most of the C16 and C18 monomers identified in the suberins also occur in plant cutins emphasizing the close chemical similarity between the two anatomical groups of lipid biopolymer.  相似文献   

2.
The suberin contents of the isolated superficial cork layers of Malus pumila stems and root ranged from 15 to 35% of the dry weight. The qualitative composition of the aliphatic monomers obtained after alkaline depolymerization of the extractive-free corks was similar but some quantitative differences were found according to cultivar and age of the cork layer. 1-Alkanols (mainly 22:0, 24:0 and 26:0), alkanoic acids (mainly 22:0 and 24:0), α, ω-alkanedioic acids (mainly 16:0, 18:1 (9) and 18:0) and ω-hydroxyalkanoic acids (mainly 18:1 (9) and 22:0) were major constituents of all the samples examined and together they comprised 40–50% of the total monomeric mixture. The remainder was composed mainly of 9,10-epoxy-18-hydroxy-and 9,10,18-trihydroxyoctadecanoic acids. The corresponding dibasic acids, 9,10-epoxy- and 9,10-dihydroxyoctadecane-1,18-dioic, were minor components as were C16 and C18 dihydroxyalkanoic acids (mainly 10,16-dihydroxyhexadecanoic and 10,18-dihydroxyoctadecanoic acids, respectively). The root suberin dittered from that of the stem in containing larger amounts of 9,10-epoxy- 18-hydroxyoctadecanoic and 18-hydroxyoctadec-9-enoic acids.  相似文献   

3.
The cuticular wax and cutin components of the cuticular membranes isolated from the leaves of two spinach cultivars have been determined. The membranes contain about 0·007 mg/cm2 of cuticular wax which comprises monobasic acids (C16–C38) with hexadecanoic as the major component. The amounts of cutin are comparable with those of cuticular wax and the monomeric constituents are predominantly C18 epoxy compounds. The most abundant monomer is 9,10-epoxy-18-hydroxyoctadecanoic acid (up to 63%) together with substantial amounts of 9,10,18-trihydroxyoctadecanoic acid (up to 22%). Also present are 9,10-epoxyoctadecane-1,18-dioic acid (6–7%) dihydroxyhexadecanoic acid (3–4%) and ω-hydroxymonobasic and fatty acid fractions. The tentative identification of two minor components, 18-hydroxyoxooctadecanoic and 9,10-epoxy-12,18-dihydroxyoctadecanoic acids, is also made. Although spinach membranes have a delicate structure their cutin composition is essentially similar to that of much more substantial membranes.  相似文献   

4.
Suberin from the roots of carrots (Daucus carota), parsnip (Pastinaca sativa), rutabaga (Brassica napobrassica), turnip (Brassica rapa), red beet (Beta vulgaris), and sweet potato (Ipomoea batatas) was isolated by a combination of chemical and enzymatic techniques. Finely powdered suberin was depolymerized with 14% BF3 in methanol, and soluble monomers (20-50% of suberin) were fractionated into phenolic (<10%) and aliphatic (13-35%) fractions. The aliphatic fractions consisted mainly of ω-hydroxyacids (29-43%), dicarboxylic acids (16-27%), fatty acids (4-18%), and fatty alcohols (3-6%). Each fraction was subjected to combined gas-liquid chromatography and mass spectrometry. Among the fatty acids very long chain acids (>C20) were the dominant components in all six plants. In the alcohol fraction C18, C20, C22, and C24 saturated primary alcohols were the major components. C16 and C18 dicarboxylic acids were the major dicarboxylic acids of the suberin of all six plants and in all cases octadec-9-ene-1, 18-dioic acid was the major component except in rutabaga where hexadecane-1, 16-dioic acid was the major dicarboxylic acid. The composition of the ω-hydroxyacid fraction was quite similar to that of the dicarboxylic acids; 18-hydroxy-octadec-9-enoic acid was the major component in all plants except rutabaga, where equal quantities of 16-hydroxyhexadecanoic acid and 18-hydroxyoctadec-9-enoic acid (42% each) were found. Compounds which would be derived from 18-hydroxyoctadec-9-enoic acid and octadec-9-ene-1, 18-dioic acid by epoxidation, and epoxidation followed by hydration of the epoxide, were also detected in most of the suberin samples. The monomer composition of the six plants showed general similarities but quite clear taxonomic differences.  相似文献   

5.
The component hydrocarbons, sterols, alcohols, monobasic, α,ω-dibasic and ω-hydroxy acids of the fresh hand decayed leaves and the pneumatophores of the mangrove Avicennia marina are reported in detail. From the quantitative comparisons which can be drawn, relative changes in the lipid classes occurring during leaf decay can be highlighted. These base-line data are important to our understanding of inputs to marine intertidal sediments. During leaf decay the only significant changes were a reduction in the total absolute concentrations of monobasic acids due largely to a decrease in concentration of the C18 polyunsaturated fatty acids, and an enhancement of the concentrations of the long-chain monobasic acids, ω-hydroxy acids and α,ω-dibasic acids. This resistance to degradation shown by the cutin derived acids (α,ω-dibasic, ω-hydroxy and long-chain monobasic acids) relative to the cellular and wax derived lipids may allow these cutin components to be used as quantitative markers of A. marina in mangrove associated sediments.  相似文献   

6.
Epicuticular wax isolated from the cotyledons and primary needles of 10-week-old Pinus radiata seedlings is similar in composition and contains 86% neutral compounds, viz. alkyl esters (25%, C24–C64), nonacosan-10-ol (52%), heptacosane-5,10-diol (2%), nonacosane-4,10-diol, nonacosane-5,10-diol, and nonacosane-10,13-diol (total 12%) and estolides, MW ca 800 (2%), MW ca 1100 (6%), and MW ca 1500 (1%). The acidic fraction (14%) contains n-acids (78%, C12–C32) and diterpene acids (22%, mainly abieta-8,11,13-trien-18-oic, with lesser amounts of pimara-8(14),15-dien-18-oic, isopimara-7,15-dien-18-oic and hydroxylated aromatic, diene and mono-ene acids). Wax isolated from primary needles of 1-yr-old seedlings had a similar neutral fraction composition, but the acidic fraction contained predominantly the diterpene acid mixture, with only trace amounts of n-acids. The wax from 1-yr-old secondary, needles from P. radiata forest trees aged 5 yr and 40 yr contained an acid fraction (12% 5 yr, 17% 40 yr trees) comprising the diterpene acid mixture, with trace amounts of n-acids together with ω-hydroxy acids (C12, C14 and C16). The neutral fraction from both young and old trees had a similar composition containing alkyl esters (7%, C24–C66), estolides (90%, MW 566-ca 1500), nonacosan-10-ol (2%) and the heptacosane and nonacosane diols (1%). During growth and maturation of P. radiata, the nonacosan-10-ol content of the needle wax decreases while the proportion of estolides and diterpene acids increases, the latter probably being located around the stomatal pore.  相似文献   

7.
Addition to Bacillus acidocaldarius of acids which can act as primers for fatty acid synthesis promote the synthesis of corresponding fatty acids competitively. The effective acids are n?C5 to -?7 (not C4 or C8), iso- and anteiso-C, and ?C, (not C4), and a range of cyclic acids from cyclobutylacetic and cyclopentanecarboxylic to cycloheptylacetic. New non-natural ω-cyclobutyl-, ω-cyclopentyl-, and ω-cycloheptyl-fatty acids are obtainable. The range of acceptable primers and the range of fatty acids produced therefrom indicate, respectively, the substrate specificities of the transacylase which introduces acyl species into fatty acids synthesis and the one which removes them. The specificity of the primer transacylase may be similar to that in some rumen anaerobes.  相似文献   

8.
The composition of the aliphatics of the protective cuticular polymers from different anatomical regions from several plant species was determined by combined gas-liquid chromatography and mass spectrometry of the depolymerization products derived from the polymers. The polymer from the aerial parts of Vicia faba showed similar composition; dihydroxypalmitic acid was the major (>85%) component of the cutin covering leaves, petioles, flower petals and stem with smaller amounts of palmitic acid and ω-hydroxy palmitic acid. On the other hand, the chief components of the polymer from the tap root were ω-hydroxy C16:0 and C18:1 acids and/or the corresponding dicarboxylic acids. The positional isomer composition of the dihydroxy C16 acids was shown to be dependent upon anatomical location, developmental stage, and light. Apple cutin from rapidly expanding organs (flower petal and stigma) was shown to contain predominately C16 family acids whereas the C18 family dominated in cutin of slower growing organs (leaf and fruit). The composition of the aliphatic components of cutin found in the seed coats of pea, corn, barley, and lettuce was found to be similar to that of the cuticular polymer of the leaves in each species.  相似文献   

9.
A Gram-positive, coccoid bacterial isolate (02-St-019/1T), forming beige pigmented colonies was obtained from an indoor air sample. Based on 16S rRNA gene sequence similarity studies it was determined that this isolate 02-St-019/1T belonged to the genus Kytococcus, showing sequence similarties of 98.6% to Kytococcus schroeteri DSM 13884T and 98.3% to Kytococcus sedentarius DSM 20547T, respectively. The diagnostic diaminoacid of the peptidoglycan was lysine, cell wall sugars were ribose and xylose. The major menaquinones detected were MK-7 and MK-8. The polar lipid profile consisted of the major phospholipids diphosphatidylglycerol, phosphatidylglycerol, phosphatidylinositol, phosphatidylserine and phosphatidylinositol mannoside. Fatty acid patterns were composed of major amounts of the iso- and anteiso-branched fatty acids anteiso C17:0, iso C15:0 and iso C17:0 and unsaturated fatty acids (C17:1 ω8c, iso C17:1 ω9c, and C17:1 ω8c) with smaller amounts of the straight-chain fatty acids C15:0, C16:0 and C17:0. The results of DNA–DNA hybridizations and physiological and biochemical tests clearly allowed a genotypic and phenotypic differentiation of strain 02-St-019/1T from the two described Kytococcus species. On the basis of these results a novel species to be named Kytococcus aerolatus sp. nov., is proposed, with the type strain 02-St-019/1T (=DSM 22179T=CCM 7639T).  相似文献   

10.
The cutins of fruits and leaves of four apple cultivars have been analysed using TLC, GLC and GC-MS. They are similarly composed of saturated, monounsaturated and diunsaturated fatty, hydroxy-fatty and epoxyhydroxy-fatty acids. The most abundant monomers are 18-hydroxyoctadeca-9,12-dienoic, 10,16-dihydroxyhexadecanoic, 9,10-epoxy-18-hydroxyoctadec-12-enoic, 9,10-epoxy-18-hydroxyoctadecanoic and 9,10,18-trihydroxyoctadecanoic acids. The fruit cutins have high contents of epoxides (35–40%) and unsaturated components ( > 40%) and C18 compounds predominate over C16. The leaf cutins contain smaller amounts of unsaturated components than the fruits and higher proportions of C16 compounds. The adaxial leaf cutin differs in composition from the abaxial. 10,16-Dihydroxyhexadecanoic and 9,10-epoxy-18-hydroxoctadecanoic acids are the major constituents (each ca. 30%) of the adaxial leaf cutin and 10,16-dihydroxyhexadecanoic acid (55–65%) predominates in the abaxial.  相似文献   

11.
The fatty acids from Grevillea robusta seed oil triglycerides contain 22.5 % ω-5 monoenes ranging in chain length from C14 to C28. C16 to C26 saturates (18 %), C18 to C24 ω-9 monoenes (55 %), C18 diene (2.3 %) and C18 triene (0.7 %) make up the remainder of the acids.  相似文献   

12.
An engineered reversal of the β-oxidation cycle was exploited to demonstrate its utility for the synthesis of medium chain (6–10-carbons) ω-hydroxyacids and dicarboxylic acids from glycerol as the only carbon source. A redesigned β-oxidation reversal facilitated the production of medium chain carboxylic acids, which were converted to ω-hydroxyacids and dicarboxylic acids by the action of an engineered ω-oxidation pathway. The selection of a key thiolase (bktB) and thioesterase (ydiI) in combination with previously established core β-oxidation reversal enzymes, as well as the development of chromosomal expression systems for the independent control of pathway enzymes, enabled the generation of C6–C10 carboxylic acids and provided a platform for vector based independent expression of ω-functionalization enzymes. Using this approach, the expression of the Pseudomonas putida alkane monooxygenase system, encoded by alkBGT, in combination with all β-oxidation reversal enzymes resulted in the production of 6-hydroxyhexanoic acid, 8-hydroxyoctanoic acid, and 10-hydroxydecanoic acid. Following identification and characterization of potential alcohol and aldehyde dehydrogenases, chnD and chnE from Acinetobacter sp. strain SE19 were expressed in conjunction with alkBGT to demonstrate the synthesis of the C6–C10 dicarboxylic acids, adipic acid, suberic acid, and sebacic acid. The potential of a β-oxidation cycle with ω-oxidation termination pathways was further demonstrated through the production of greater than 0.8 g/L C6–C10 ω-hydroxyacids or about 0.5 g/L dicarboxylic acids of the same chain lengths from glycerol (an unrelated carbon source) using minimal media.  相似文献   

13.
Cutin and suberin polymers from various anatomical regions of grapefruit were analyzed chemically and ultrastructurally. The leaf, fruit peel and juice-sac showed an amorphous cuticular layer. The cutin in the leaf was composed of 10,16-dihydroxy C16 acid and its positional isomers as the major monomers whereas 16-hydroxy-10-oxo C16 acid was a major component in the fruit peel. Juice-sac cutin, on the other hand, contained the dihydroxy C16 acids, hydroxyoxo C16 acids, hydroxyepoxy C18 acids and trihydroxy C18 acids. Ultrastructural examination of the inner seed coat showed that an amorphous cuticular layer encircled the entire seed except in the chalazal region which showed several layers of cells with lamellar suberin structure throughout the cell walls. Consistent with the ultrastructural assignment, the compositions of the aliphatic components of the polymers from the chalazal region and the non-chalazal region indicated the presence of suberin and cutin, respectively. The aliphatic portion of the polymer from the chalazal region of the inner seed coat contained C16, C18:1, C22 and C24 -hydroxy acids (46% combined total) and the corresponding dicarboxylic acids (43%) as the major components. -Hydroxy-9,10-epoxy C18 acids and 9,10,18-trihydroxy C18 acids were the major components (77%) of the polymer from the non-chalazal portion of the inner seed coat. The main portion and the chalazal region of the inner seed coat yielded 17 and 342 g/cm2 of aliphatic monomers, respectively, and the diffusion resistance of these two portions of the inner seed coat were 62 and 192 sec/cm, respectively. The inner seed coat was shown to be the major moisture diffusion barrier influencing imbibition and germination.Scientific Paper No. 5649, Project 2001, College of Agriculture Research Center, Washington State University, Pullman, Washington 99164  相似文献   

14.
Recent investigations have shown macromolecules, such as cutins, and suberins as effective markers for above and belowground plant tissues. These biopolyesters contain structural units specific for different litter components and for root biomass. The aim of this work was to understand the fate of plant organic matter (OM) in Mediterranean forest soils by evaluating the incorporation of cutin and suberin by measuring specific biomarkers. Soil and plant tissue (leaves, woods and roots) samples were collected in two mixed Mediterranean forests of Quercus ilex (holm oak) in costal stands in Tuscany (central Italy), which have different ecological and edaphic features. Ester-bound lipids of mineral and organic horizons and the overlying vegetation were analysed using the saponification method in order to depolymerise cutins and suberins and release their specific structural units. Cutin and suberin specific aliphatic monomers were identified and quantified by gas chromatographic techniques. The distribution of cutin and suberin specific monomers in plant tissue suggested that mid-chain hydroxy acids can be used as leaf-specific markers and α,ω-alkanedioic acids and ωC18:1 as root-specific markers. Differences in the distributions of biomarkers specific for above and belowground plant-derived OM was observed in the two types of soils, suggesting contrasted degradation, stabilisation and transport mechanisms that may be related to soil physico-chemical properties. The acidic and dry soil appeared to inhibit microbial activity, favouring stabilization of leaf-derived compounds, while, in the more fertile soil, protection within aggregates appeared to better preserve root-derived compounds.  相似文献   

15.
Three C18 epoxy acids occur in plant cutins and suberins. 9,10-Epoxy-18-hydroxyoctadecanoic acid is a common constituent of both cutins and suberins whilst 9,10-epoxy-18-hydroxyoctadec-12-enoic acid is also present in some cutins. 9,10-Epoxyoctadecane-1,18-dioic acid occurs more commonly in suberins. Epoxy acids may comprise up to 60% of the total monomers obtained from some polymers. The epoxy compounds are readily converted into their corresponding alkoxyhydrin alkyl esters on depolymerization of cutin or suberin by alcoholysis. The chromatographic and MS properties of the alkoxyhydrin derivatives enable them to be readily distinguished from other cutin and suberin hydroxyfatty acids and to be used for the qualitative and quantitative determination of epoxy acids in the polymers.  相似文献   

16.
Cold-loving microorganisms developed numerous adaptation mechanisms allowing them to survive in extremely cold habitats, such as adaptation of the cell membrane. The focus of this study was on the membrane fatty acids of Antarctic Flavobacterium spp., and their adaptation response to cold-stress. Fatty acids and cold-response of Antarctic flavobacteria was also compared to mesophilic and thermophilic members of the genus Flavobacterium. The results showed that the psychrophiles produced more types of major fatty acids than meso- and thermophilic members of this genus, namely C15:1 iso G, C15:0 iso, C15:0 anteiso, C15:1 ω6c, C15:0 iso 3OH, C17:1 ω6c, C16:0 iso 3OH and C17:0 iso 3OH, summed features 3 (C16:1 ω7cand/or C16:1 ω6c) and 9 (C16:0 10-methyl and/or C17:1 iso ω9c). It was shown that the cell membrane of psychrophiles was composed mainly of branched and unsaturated fatty acids. The results also implied that Antarctic flavobacteria mainly used two mechanisms of membrane fluidity alteration in their cold-adaptive response. The first mechanism was based on unsaturation of fatty acids, and the second mechanism on de novo synthesis of branched fatty acids. The alteration of the cell membrane was shown to be similar for all thermotypes of members of the genus Flavobacterium.  相似文献   

17.
Cutin, the structural component of plant cuticle, is a polymer of C16 and C18 hydroxy fatty acids. Previous results have suggested that oleic acid undergoes ω-hydroxylation, epoxidation of the double bond, and, finally, hydration of the epoxide to give rise to the three major components of the C18 family of cutin acids. 18-Hydroxy [18-3H]oleic acid and 18-hydroxy-9,10-epoxy[18-3H]stfaric acid have been synthesized and, with these synthetic substrates, the conversion of 18-hydroxyoleic acid to 18-hydroxy-9,10-epoxystearic acid and the hydrolysis of 18-hydroxy-9,10-epoxystearic acid to 9,10,18-trihydroxystearic acid were directly demonstrated in apple fruit skin and in the leaves of apple and Senecio odoris. Trichloropropene oxide, an inhibitor of microsomal epoxide hydrases of animals, specifically inhibited the conversion of [1-14C]oleic acid into 18-hydroxy-9,10-epoxystearic acid and 9,10,18-trihydroxystearic acid, while it had no effect on the conversion of [1-14C]palmitic acid into hydroxylated palmitic acid, a process which does not involve epoxy acid intermediates. Therefore, it appears that this inhibitor affects epoxidation and or epoxide hydration steps involved in cutin biosynthesis.  相似文献   

18.
Changes in response to temperature of lipid classes, fatty acid composition and mRNA levels for acyl-lipid desaturase genes were studied in the marine unicellular cyanobacterium, Synechococcus sp. PCC 7002. The degree of unsaturation of C18 fatty acids increased in cells grown at lower temperature for all lipid classes, and ω3 desaturation occurred specifically in cells grown at low temperature. While the level of 18:1(9) fatty acids declined, desaturation at the ω3 position of C18 fatty acids increased gradually during a 12-h period after a temperature shift-down to 22°C. However, the mRNA levels of the desA (Δ12 desaturase), desB (ω3 desaturase) and desC (Δ9 desaturase) genes increased within 15 min after a temperature shift-down to 22°C; the desaturase gene mRNA levels also rapidly declined within 15 min after a temperature shift-up to 38°C. Therefore, the elevation of mRNA levels for the desaturase genes is not the rate-limiting event for the increased desaturation of membrane lipids after a temperature shift-down. The rapid, low-temperature-induced changes in mRNA levels occurred even when cells were grown under light-limiting conditions for which the growth rates at 22°C and 38°C were identical. These studies indicate that the ambient growth temperature, and not some other growth rate-related process, regulates the expression of acyl lipid desaturation in this cyanobacterium.  相似文献   

19.
Phospholipid and acid composition of 5 strains of ‘true’ Nocardia and 4 strains of nocardoid bacteria have been studied. A great homogeneity was found in all the Nocardia species: phospholipids consist of cardiolipin, phosphatidyl ethanolamine, phosphatidylinositol and phosphatidylinositol mannoside. Streptomyces (Nocardia) mediterranei did not contain phosphatidylinositol and Oerskovia (Nocardia) turbata had no phosphatidyl ethanolamine. The fatty acid composition of these phospholipids was determined and was found different in Nocardia and nocardoid species. Nocardia were rich in straight chain fatty acids and tuberculostearic acid while the phospholipids of nocardoid bacteria contained greater amounts of branched fatty acids. The fatty acids from acetone soluble lipids consisted of hydroxy and non-hydroxy compounds. Hydroxy acids were found in Nocardia which contained nocardic acids: high MW β-hydroxy α-branched acids and in S. mediterranei which contained β-hydroxy acids with 15–17 carbon atoms. Non-hydroxy acids were essentially palmitic and tuberculostearic acids in Nocardia species while S. mediterranei and O. turbata contained great amounts of iso acids from C14 to C17. Phospholipid and acid composition are discussed as criteria of taxonomic classification of Nocardia and related Actinomycetes.  相似文献   

20.
Wax on leaves of Agropyron intermedium contains hydrocarbons (11%, C27–C33), esters (11%, C32–C60), free alcohols (180%, C26) 25-oxohentriacontane-14,16-dione (17%), 10-oxohentriacontane-14,16-dione (5y%), 25-hydroxyhentriacontane-14,16-dione (12%) and 26-hydroxyhentriacontane-14,16-dione (2%). Wax on spikes contains additional components, C25–C33cis 9-alkenes (32% of hydrocarbons), and more β-diketones, 25-hydroxy (17%) and 26-hydroxy (3%) hentriacontane-14,16-diones, 10,25-dioxohentriacontane-14,16-dione (1%) and 4-hydroxy-25-oxo-(2%), 25-hydroxy-10-oxo-(1.3%) and 26-hydroxy-10-oxo-(0.7%) hentriacontane-14,16-diones; free alcohols were very minor components (1%, C24–C32).  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