首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Summary Four cathodal bands (C1, C2, C3 and C4) of esterase (E1, C1. 3.1) were correlated with the grain weight of rice (Oryza sativa L.). Zymogram patterns indicated intensity differences among these bands infinegrain and coarse-grain varieties. Bands C1. and C2 were dark in fine grain varieties whereas C3 and C4 were dark in coarse grain varieties. These bands were specific to endosperm. Observations on fine-grain (Kalanamak), coarse grain (SR(26)B) varieties and their reciprocal hybrids indicated the presence of 4 esterase loci G1, G2, G3 and G4, corresponding to bands C1, C2, C3 and C4, respectively. A possible model for heterosis in grain weight of rice was proposed which supports the dominance theory of heterosis. In hybrid vigour the 4 esterase loci appear to be associated with grain weight and they complemented each other in an additive manner.  相似文献   

2.
The molecular conformation of the monoclinic crystalline polymorph of prostaglandin A1 has been determined by X-ray diffraction techniques. The space group is P21 with a = 13.637 (2), b = 7.567 (1), I c = 10.576 (2) Å, β = 107.37 (3)°; Dc = 1.073 g·cm−3 for Z = 2. The molecular conformation is characterized by the nearly parallel arrangement of the C1–C7 and C13–C20 side chains, with a general flattening of the overall structure when compared with the orthorhombic polymorph. The cyclopentenone moiety assumes a C8 envelope conformation with C8 and O9 displaced +0.29 Å and −0.18 Å from the C9–C10=C11–C12 plane respectively. Concerted, small variations of the torsion angles, primarily about the C8–C12, C14–C15 and C16–C17 bonds, bring the monoclinic and orthorhombic conformations into coincidence.  相似文献   

3.
Two Bacillus strains were isolated from the foregut of the water beetle Agabus affinis (Payk.) and tested for their steroid transforming ability. After incubation with androst-4-en-3,17-dione (AD), 13 different transformation products were detected. AD was hydroxylated at C6, C7, C11 and C14, resulting in formation of 6β-, 7α-, 11α- and 14α-hydroxy-AD. One strain also produced small amounts of 6β,14α-dihydroxy-AD. Partly, the 6β-hydroxy group was further oxidized to the corresponding 6-oxo steroids. In addition, a specific reduction of the Δ4-double bond was observed, leading to the formation of 5α-androstane derivatives. In minor yields the carbonyl functions at C3 and C17 were reduced leading to the formation of 3ξ-OH or 17β-OH steroids. EI mass spectra of the trimethylsilyl and O-methyloxime trimethylsilyl ether derivatives of some transformation products are presented for the first time.  相似文献   

4.
Summary Analyses of carbon-assimilation patterns in response to intercellular CO2 concentrations, and the photosynthetic water-and nitrogen-use efficiencies, were conducted for a C3, a C4, and three C3–C4 species in the genus Flaveria in order to determine some of the advantages and disadvantages of C3–C4 intermediate photosynthesis. Operational intercellular CO2 partial pressures (pi), determined when the atmospheric CO2 partial pressure (pa) was approximately 330 bar, in the C3–C4 species were generally equal to, or greater than, those observed in the C3 species under well-watered or water-stressed conditions. This reflects equal, or lower, water-use efficiencies (WUEs) in the C3–C4 species. The only case in which higher WUEs were observed in the C3–C4 species, compared to the C3 species, was when photosynthesis rates were limited by available nitrogen and were less than 12.5 mol CO2 m-2s-1. At higher photosynthesis rates, the C3–C4 species exhibited lower values of photosynthesis rate for equal values of stomatal conductance (lower WUE), compared to the C3 species. Comparing slopes for the linear regions of the relationship between leaf nitrogen content and net photosynthesis rate (taken as an index of photosynthetic nitrogen-use efficiency, NUE), the C4 species exhibited the highest NUE, followed by the C3–C4 species, F. ramosissima, with the other two C3–C4 species and the C3 species being equal and exhibiting the lowest NUEs. The lack of consistent advantages in NUE and WUE in the C3–C4 species F. pubescens and F. floridana suggest that in some C3–C4 Flaveria species C4-like anatomy and biochemistry do not provide the same gas exchange advantages that we typically attribute to the CO2-concentrating mechanism of fully-expressed C4 plants.  相似文献   

5.
Previous work from our laboratory has shown dinoflagellates, which possess the carotenoid peridinin, have been divided into two clusters based on plastid galactolipid fatty acid composition. In one cluster major forms of monogalactosyldiacylglycerol (MGDG) and digalactosyldiacylglycerol (DGDG), lipids that comprise the majority of photosynthetic membranes, were C18/C18 (sn‐1/sn‐2), with octadecapentaenoic [18:5(n‐3)] and octadecatetraenoic [18:4(n‐3)] acid as principal fatty acids. The other cluster contained C20/C18 major forms, with eicosapentaenoic acid [20:5(n‐3)] being the predominant sn‐1 fatty acid. In this study, we have found that Symbiodinium microadriaticum isolated from the jellyfish, Cassiopea xamachana, when grown at 30°C, produced MGDG and DGDG with a more saturated fatty acid, 18:4(n‐3), at the sn‐2 carbon than when grown at 20°C where 18:5(n‐3) predominates. This modulation of the sn‐2 fatty acid's level of saturation is mechanistically similar to what has been observed in Pyrocystis, a C20/C18 dinoflagellate. We have also examined the effect of growth temperature on the betaine lipid, diacylglycerylcarboxyhydroxymethylcholine (DGCC), which has been observed by others to be the predominant non plastidial polar lipid in dinoflagellates. Temperature effects on it were minimal, with very few modulations in fatty acid unsaturation as observed in MGDG and DGDG. Rather, the primary difference seen at the two growth temperatures was the alteration of the amount of minor forms of DGCC, as well as a second betaine lipid, diacylglyceryl‐N,N,N‐trimethylhomoserine.  相似文献   

6.
Photon requirements for O2-evolution in red (λ=680nm) light (Фr) were measured for six C3 species, one C3-like, C3–C4 intermediate species, and three C4 species, including examples of NADP-malic enzyme and PEP-carboxykinase C4 sub-groups. Variation in Фr within the C3 species was small with a mean value of 7.96 ±0.12 mol photon mol−1 O2, whereas the mean value for the C4 species was 12.27± 1.53 mol photon mol−1 O2, with the lowest value, 9.24 ±0.13 mol photon mol−1 O2, for the PEP-carboxykinase C4 species Spartina townsendii. The C3–C4 intermediate species Panicum milioides had a value of 9.05 ±0.29 mol photon mol−1 O2, approximately 1 mol photon mol−1 O2 greater than the C3 species. The possibility that this extra cost is due to PEP-carboxylase-dependent recycling of CO2 is discussed. No correlation was found between Фr and chlorophyll content or leaf absorptance. Based on white (ФW) and red light measurements of the photon requirement, values in red light were approximately 20% higher than white-light estimates. These results are discussed with reference to accepted mechanisms of energy transduction in thylakoid membranes (Z-scheme), expected inefficiencies and losses during light-harvesting and electron transport reactions, and the influence of respiratory processes.  相似文献   

7.
Photosynthetic pathways (C3, C4, and CAM) and morphological functional types were identified for the species from vegetation in agro-pastoral ecotone, North Beijing. 792 vascular plant species (nearly half of the total species in the ecotone), in 66 families and 317 genera, were identified with C3, C4, and CAM photosynthesis (Table 1). 710 species (90 % of the identified species in Table 1) in 268 genera and 64 families were found with C3 photosynthesis, 68 species (9 % of the total identified species) in 40 genera and 7 families with C4 photosynthesis, and 14 species in 4 genera and 1 family with CAM photosynthesis. Gramineae is the leading family with C4 photosynthesis (43 species), Cyperaceae ranks the second (16 species) followed by Chenopodiaceae (5 species). The significant increase of C4 proportion (C4/total species) with land deterioration suggested the plants of this type are remarkably responsive to land use in the ecotone. 792 species were classified into nine morphological functional types and the changes of most of these types (e.g. perennial forbs (PEF), annual grasses (ANG), and annual forbs (ANF)) were consistent with habitats and vegetation dynamics in the agro-pastoral ecotone. Hence the photosynthetic pathways, combined with the morphological functional types, are efficient indications for studying the linkage between species and ecosystems in the ecotone.  相似文献   

8.
We determined the interactive effects of irradiance, elevated CO2 concentration (EC), and temperature in carrot (Daucus carota var. sativus). Plants of the cv. Red Core Chantenay (RCC) were grown in a controlled environmental plant growth room and exposed to 3 levels of photosynthetically active radiation (PAR) (400, 800, 1 200 μmol m−2 s−1), 3 leaf chamber temperatures (15, 20, 30 °C), and 2 external CO2 concentrations (C a), AC and EC (350 and 750 μmol mol−1, respectively). Rates of net photosynthesis (P N) and transpiration (E) and stomatal conductance (g s ) were measured, along with water use efficiency (WUE) and ratio of internal and external CO2 concentrations (C i/C a). P N revealed an interactive effect between PAR and C a. As PAR increased so did P N under both C a regimes. The g s showed no interactive effects between the three parameters but had singular effects of temperature and PAR. E was strongly influenced by the combination of PAR and temperature. WUE was interactively affected by all three parameters. Maximum WUE occurred at 15 °C and 1 200 μmol m−2 s− 1 PAR under EC. The C i /C a was influenced independently by temperature and C a. Hence photosynthetic responses are interactively affected by changes in irradiance, external CO2 concentration, and temperature. EC significantly compensates the inhibitory effects of high temperature and irradiance on P N and WUE.  相似文献   

9.
Summary We tested the hypothesis that C4 grasses are inferior to C3 grasses as host plants for herbivorous insects by measuring the relative performance of larvae of a graminivorous lepidopteran, Paratrytone melane (Hesperiidae), fed C3 and C4 grasses. Relative growth rates and final weights were higher in larvae fed a C3 grass in Experiment I. However, in two additional experiments, relative growth rates and final weights were not significantly different in larvae fed C3 and C4 grasses. We examined two factors which are believed to cause C4 grasses to be of lower nutritional value than C3 grasses: foliar nutrient levels and nutrient digestibility. In general, foliar nutrient levels were higher in C3 grasses. In Experiment I, protein and soluble carbohydrates were digested from a C3 and a C4 grass with equivalent efficiencies. Therefore, differences in larval performance are best explained by higher nutrient levels in the C3 grass in this experiment. In Experiment II, soluble carbohydrates were digested with similar efficiencies from C3 and C4 grasses but protein was digested with greater efficiency from the C3 grasses. We conclude (1) that the bundle sheath anatomy of C4 grasses is not a barrier to soluble carbohydrate digestion and does not have a nutritionally significant effect on protein digestion and (2) that P. melane may consume C4 grasses at compensatory rates.  相似文献   

10.
Tetrachloroethene (C2Cl4) dechlorination kinetics in upflow anaerobic sludge blanket (UASB) reactors was determined after introducing de novo activities into the granular sludge. These activities were introduced by immobilizing Dehalospirillum multivorans in a test reactor containing unsterile granular sludge, and in a reference reactor, R1, containing sterile granular sludge. A second reference reactor, R2, contained only unsterile granular sludge and served as a control. The kinetic experiments were performed by pulsing the reactors with C2Cl4 in a recirculating batch mode. Formate and acetate were added as electron donor and carbon source. Both reactors inoculated with D. multivorans dechlorinated C2Cl4 to an equimolar amount of C2H2Cl2 with only traces of C2HCl3 in the effluent. In the control reactor, C2HCl3 accumulated before C2H2Cl2 was produced. A computer simulation program (AQUASIM) was used to estimate the kinetic parameters. The half-saturation constants (K s) for C2Cl4 and C2HCl3 were almost equal in the reactors containing D.␣multivorans (17 μM and 18 μM for C2Cl4; 26 μM and 28 μM for C2HCl3), indicating no influence of sludge bacteria on the affinity of D. multivorans for C2Cl4 and C2HCl3. The maximum dechlorination rates (k m X B) were about twice as high in the reactor containing D.␣multivorans immobilized in sterile sludge (11 mmol C2Cl4 l sludge−1 day−1 and 27 mmol C2HCl3 l sludge−1 day−1) than in the test reactor (4.4 mmol C2Cl4 l sludge−1 day−1 and 15 mmol C2HCl3 l sludge−1 day−1). Compared to other C2Cl4-degrading systems, the dechlorination rates of the inoculated reactors and their affinities for C2Cl4 and C2HCl3 were high. Therefore, introduction of de novo activity is promising for the use of anaerobic reactors to bioremediate C2Cl4-polluted water. Received: 5 November 1998 / Received revision: 25 January 1999 / Accepted: 31 January 1999  相似文献   

11.
Evidence is presented contrary to the suggestion that C4 plants grow larger at elevated CO2 because the C4 pathway of young C4 leaves has C3-like characteristics, making their photosynthesis O2 sensitive and responsive to high CO2. We combined PAM fluorescence with gas exchange measurements to examine the O2 dependence of photosynthesis in young and mature leaves of Panicum antidotale (C4, NADP-ME) and P. coloratum (C4, NAD-ME), at an intercellular CO2 concentration of 5 Pa. P. laxum (C3) was used for comparison. The young C4 leaves had CO2 and light response curves typical of C4 photosynthesis. When the O2 concentration was gradually increased between 2 and 40%, CO2 assimilation rates (A) of both mature and young C4 leaves were little affected, while the ratio of the quantum yield of photosystem II to that of CO2 assimilation (ΦPSII/ΦCO2) increased more in young (up to 31%) than mature (up to 10%) C4 leaves. A of C3 leaves decreased by 1·3 and ΦPSII/ΦCO2 increased by 9-fold, over the same range of O2 concentrations. Larger increases in electron transport requirements in young, relative to mature, C4 leaves at low CO2 are indicative of greater O2 sensitivity of photorespiration. Photosynthesis modelling showed that young C4 leaves have lower bundle sheath CO2 concentration, brought about by higher bundle sheath conductance relative to the activity of the C4 and C3 cycles and/or lower ratio of activities of the C4 to C3 cycles.  相似文献   

12.
The rapid increase in atmospheric CO2 concentrations (Ca) has resulted in extensive research efforts to understand its impact on terrestrial ecosystems, especially carbon balance. Despite these efforts, there are relatively few data comparing net ecosystem exchange of CO2 between the atmosphere and the biosphere (NEE), under both ambient and elevated Ca. Here we report data on annual sums of CO2 (NEEnet) for 19 years on a Chesapeake Bay tidal wetland for Scirpus olneyi (C3 photosynthetic pathway)‐ and Spartina patens (C4 photosynthetic pathway)‐dominated high marsh communities exposed to ambient and elevated Ca (ambient + 340 ppm). Our objectives were to (i) quantify effects of elevated Ca on seasonally integrated CO2 assimilation (NEEnet = NEEday + NEEnight, kg C m?2 y?1) for the two communities; and (ii) quantify effects of altered canopy N content on ecosystem photosynthesis and respiration. Across all years, NEEnet averaged 1.9 kg m?2 y?1 in ambient Ca and 2.5 kg m?2 y?1 in elevated Ca, for the C3‐dominated community. Similarly, elevated Ca significantly (P < 0.01) increased carbon uptake in the C4‐dominated community, as NEEnet averaged 1.5 kg m?2 y?1 in ambient Ca and 1.7 kg m?2 y?1 in elevated Ca. This resulted in an average CO2 stimulation of 32% and 13% of seasonally integrated NEEnet for the C3‐ and C4‐dominated communities, respectively. Increased NEEday was correlated with increased efficiencies of light and nitrogen use for net carbon assimilation under elevated Ca, while decreased NEEnight was associated with lower canopy nitrogen content. These results suggest that rising Ca may increase carbon assimilation in both C3‐ and C4‐dominated wetland communities. The challenge remains to identify the fate of the assimilated carbon.  相似文献   

13.
The crystal and molecular structures of two α-aminoisobutyric acid (Aib)-containing diketopiperazines, cyclo(Aib-Aib) 1 and cyclo(Aib-L -Ile) 2 , are reported. Cyclo(Aib-Aib) crystallizes in the space group P1 with a = 5.649(3), b = 5.865(2), c = 8.363(1), α = 69.89(6), β = 113.04(8), γ = 116.0(3), and Z = 1, while 2 occurs in the space group P212121 with a = 6.177(1), b = 10.791(1), c = 16.676(1), and Z = 4. The structures of 1 and 2 have been refined to final R factors of 0.085 and 0.086, respectively. In both structures the diketopiperazine ring shows small but significant deviation from planarity. A very flat chair conformation is adopted by 1, in which the Cα atoms are displaced by 0.07 Å on each side of the mean plane, passing through the other four atoms of the ring. Cyclo(Aib-Ile) favors a slight boat conformation, with Aib Cα and Ile Cα atoms displaced by 0.11 and 0.05 Å on the same side of the mean plane formed by the other ring atoms. Structural features in these two molecules are compared with other related diketopiperazines.  相似文献   

14.
15.
The use of stable carbon isotope analysis in rooting studies   总被引:1,自引:0,他引:1  
Summary Stable carbon isotope analysis was evaluated as a means of predicting the relative proportions of C3 and C4 root phytomass in species mixtures. The following mixtures of C3 and C4 species were used: 1) big bluestem (Andropogon gerardii)/cheatgrass (Bromus tectorum), 2) little bluestem (Schizachyrium scoparium)/cheatgrass, and 3) sorghum (Sorghum bicolor)/sunflower (Helianthus annuus). There was a significant correlation (P<0.01) between % C4 phytomass and stable carbon isotope values for each of the three combinations (r 2>0.98). Root length per mass varied among the five species studied (10.1–94.3 m/g), which resulted in different conclusions depending on whether root values are expressed as length or mass. For example, field samples from a tallgrass prairie site were estimated to contain about 20% cheatgrass on a mass basis, whereas the figure was 68% when expressed in terms of length. The combination of stable carbon isotope analysis with length-for-mass measurements promises to be a useful means of studying root competition between C3 and C4 plants.  相似文献   

16.
Soil microbial biomass C (Cmic) is a sensitive indicator of trends in organic matter dynamics in terrestrial ecosystems. This study was conducted to determine the effects of tropospheric CO2 or O3 enrichments and moisture variations on total soil organic C (Corg), mineralizable C fraction (CMin), Cmic, maintenance respiratory (qCO2) or Cmic death (qD) quotients, and their relationship with basal respiration (BR) rates and field respiration (FR) fluxes in wheat‐soybean agroecosystems. Wheat (Triticum aestivum L.) and soybean (Glycine max. L. Merr) plants were grown to maturity in 3‐m dia open‐top field chambers and exposed to charcoal‐filtered (CF) air at 350 μL CO2 L?1; CF air + 150 μL CO2 L?1; nonfiltered (NF) air + 35 nL O3 L?1; and NF air + 35 nL O3 L?1 + 150 μL CO2 L?1 at optimum (? 0.05 MPa) and restricted soil moisture (? 1.0 ± 0.05 MPa) regimes. The + 150 μL CO2 L?1 additions were 18 h d?1 and the + 35 nL O3 L?1 treatments were 7 h d?1 from April until late October. While Corg did not vary consistently, CMin, Cmic and Cmic fractions increased in soils under tropospheric CO2 enrichment (500 μL CO2 L?1) and decreased under high O3 exposures (55 ± 6 nL O3 L?1 for wheat; 60 ± 5 nL O3 L?1 for soybean) compared to the CF treatments (25 ± 5 nL O3 L?1). The qCO2 or qD quotients of Cmic were also significantly decreased in soils under high CO2 but increased under high O3 exposures compared to the CF control. The BR rates did not vary consistently but they were higher in well‐watered soils. The FR fluxes were lower under high O3 exposures compared to soils under the CF control. An increase in Cmic or Cmic fractions and decrease in qCO2 or qD observed under high CO2 treatment suggest that these soils were acting as C sinks whereas, reductions in Cmic or Cmic fractions and increase in qCO2 or qD in soils under elevated tropospheric O3 exposures suggest the soils were serving as a source of CO2.  相似文献   

17.
Ethanolamine plasmalogens (1-alk-1′-enyl-2-acyl-sn-glycero-3-phosphoethanolamines) of many tissues contain high levels of arachidonate at their 2-position, and in certain tissues have been implicated as possible donors of arachidonate required in the synthesis of prostaglandins and thromboxanes. In the present study, [3H]arachidonate-labeled phospholipids of HSDM1C1 cells, a cell line derived from a mouse fibrosarcoma, were examined to determine the donor of the arachidonic acid released upon bradykinin stimulation of the synthesis of PGE2. HSDM1C1 cells labeled with [3H]arachidonic acid for 24 hr in serum-free medium were used in most of the experiments and had the following distribution of label among the cellular lipids; phosphatidylcholine (33%), phosphatidylinositol (20%), diacyl-sn-glycero-3-phosphoethanolamine (15%), ethanolamine plasmalogen (15%), and less polar lipids (16%). Bradykinin treatment stimulated a rapid hydrolysis of [3H]arachidonate from the cellular lipids and conversion of the released acid to PGE2, which was secreted into the medium. The label was released predominantly from phosphatidylinositol and possibly from phosphatidylcholine with no detectable change in the labeling of diacyl- or 1-alk-1′-enyl-2-acyl-sn-glycero-3-phosphoethanolamine. The ethanolamine plasmalogens, therefore, do not appear to be involved in the stimulated release of arachidonate in the HSDM1C1 cells. Indomethacin blocked the bradykinin-stimulated synthesis of PGE2 and to a lesser degree inhibited the release of [3H]-arachidonate from the cellular lipids into the medium.  相似文献   

18.

Background  

The key enzymes of photosynthetic carbon assimilation in C4 plants have evolved independently several times from C3 isoforms that were present in the C3 ancestral species. The C4 isoform of phosphoenolpyruvate carboxylase (PEPC), the primary CO2-fixing enzyme of the C4 cycle, is specifically expressed at high levels in mesophyll cells of the leaves of C4 species. We are interested in understanding the molecular changes that are responsible for the evolution of this C4-characteristic PEPC expression pattern, and we are using the genus Flaveria (Asteraceae) as a model system. It is known that cis-regulatory sequences for mesophyll-specific expression of the ppcA1 gene of F. trinervia (C4) are located within a distal promoter region (DR).  相似文献   

19.
Variable factors affecting the enzymatic isolation of mesophyll protoplasts from Triticum aestivum (wheat), a C3 gras, and mesophyll protoplasts and bundle sheath strands from Digitaria sanguinalis (crabgrass), a C4 grass, have been examined with respect to yields and also photosynthetic capacity after isolation. Preparations with high yields and high photosynthetic capacity were obtained when small transverse leaf segments were incubated in enzyme medium in the light at 30°C, without mechanical shaking and without prior vacuum infiltration. Best results were obtained with an enzyme medium that included 0.5 M sorbitol, 1 mM MgCl2, 1 mM KH2PO4, 2% cellulase and 0.1% pectinase at pH 5.5. In gerneral, leaf age and leaf segment size were important factors, with highest yields and photosynthetic capacities obtained from young leaves cut into segments less than 0.8 mm. To facilitate the cutting of such small segments, a mechanical leaf cutter is described that uniformly (± 0.05 mm) cuts leaf tissue into transverse segments of variable size (0.4–2 mm). Isolations that required more than roughly 4 h gave poor yields with reduced photosynthetic capacity; however, using the optimum conditions described, functional preparations could be roughly 2 h. High rates of light dependent CO2 fixation by the C4 mesophyll protoplasts required the addition of pyruvate and low levels of oxalacetate, while isolated bundle sheath strands and C3 mesophyll protoplasts supported CO2 fixation without added substrates. Rates of CO2 fixation by isolated wheat protoplasts generally exceeded the reported rates of whole leaf photosynthesis. Wheat mesophyll protoplasts and crabgrass bundle sheath strands were stable when stored at 4°C while C4 mesophyll protoplasts were stable when stored at 25°C.  相似文献   

20.
Three racemic butanolides, majorenolide ( 1 ), majorynolide ( 2 ), and majoranolide ( 3 ), with 18 known compounds, including ten butanolides, i.e., litsenolide A2 ( 4 ), litsenolide B2 ( 5 ), litsenolide C1 ( 6 ), litsenolide C2 ( 7 ), hamabiwalactone A ( 8 ), hamabiwalactone B ( 9 ), litseakolide A ( 10 ), litseakolide B ( 11 ), isoobtusilactone ( 12 ), and obtusilactone ( 13 ); one lignan, i.e., (±)‐syringaresinol ( 14 ), two flavans, i.e., (+)‐catechin ( 15 ), and (?)‐epicatechin ( 16 ), one coumarin, i.e., scopoletin ( 17 ), and four steroids, i.e., a mixture of β‐sitosterol ( 18 ) and stigmasterol ( 19 ), and a mixture of β‐sitosteryl‐3‐O‐β‐D ‐glucoside ( 20 ) and stigmasteryl‐3‐O‐β‐D ‐glucoside ( 21 ) were isolated from the root of Lindera akoensis. The structures of the isolates were elucidated by in‐depth spectroscopic analysis. Compounds 1 – 3 were previously assigned a δ‐lactone structure, which was then revised to a γ‐lactone structure, based on 1D‐NMR data. The cigar‐HMBC technique was used to confirm the accuracy of the γ‐lactone structure, and the zero [α] value of compounds 1 – 3 suggested that they were considerably racemized. Nine butanolides 1 – 3, 4 – 8 , and 10 showed antimycobacterial activities against M. tuberculosis H37Rv, with MIC values of 15–50 μg/ml.  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