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 109 毫秒
1.
The geographical distribution and analysis of the essential oils of species from three sections of Hypericum L. (Guttiferae/Clusiaceae/Hypericaceae) from Portugal are presented. Hypericum perfoliatum (section Drosocarpium) grows wild in the centre and south of Portugal; Hypericum humifusum and Hypericum linarifolium are both from section Oligostema, the former occurring throughout the country, while the second is distributed mainly in the north and centre; Hypericum pulchrum (section Taeniocarpium) is confined to the littoral north of Portugal. The essential oils were obtained by distillation–extraction, hydrodistillation and distillation in a modified Marcusson apparatus from the dried aerial parts of the different populations and were analysed by GC and GC–MS. Monoterpene hydrocarbons constituted the main fraction in all oils (43–69%, 53–85%, 28–45% and 48–65% for H. perfoliatum, H. humifusum, H. linarifolium and H. pulchrum, respectively). Sesquiterpene hydrocarbons (2–13%, 6–18%, 21–27% and 16–18%, respectively) and a third fraction of non-terpenic compounds (20–29%, 3–16%, 2–14% and 5–11%, respectively) from the four species attained relatively high amounts in all oils. Within each species, no major differences were detected in the essential oil composition, despite the fact that different locations, phenological phases and extraction methodologies were used. Notwithstanding the dominance of α-pinene in all four species' oils, cluster and principal components analysis on the identified components showed that the range of α-pinene, β-pinene and n-nonane supported a separation of the four species. The essential oil composition of the four species showed some qualitative resemblances, which correlate well with the taxonomical classification based on morphological characters.  相似文献   

2.
The present study reports the isolation of sixteen lignans (1–16), an aporphine alkaloid (17), two phenolic amides (18 and 19), and a germacrane sesquiterpene lactone (20) from T. odorum. Except for compound 20, the other isolated constituents are firstly reported from T. odorum. And in this study, three polylignans (11–13) and six norlignans (5–7, 14–16) are reported for the first time from the Magnoliaceae family, which may support the opinion to establish Tsoongiodendron as a monotypic genus from a chemotaxonomical point of view.  相似文献   

3.
Strains VGXO14T and Vi1 were isolated from the Atlantic intertidal shore from Galicia, Spain, after the Prestige oil spill. Both strains were Gram-negative rod-shaped bacteria with one polar inserted flagellum, strictly aerobic, and able to grow at 18–37 °C, pH 6–10 and 2–10% NaCl. A preliminary analysis of the 16S rRNA and the partial rpoD gene sequences indicated that these strains belonged to the Pseudomonas genus but were distinct from any known Pseudomonas species. A polyphasic taxonomic approach including phylogenetic, chemotaxonomic, phenotypic and genotypic data confirmed that the strains belonged to the Pseudomonas pertucinogena group. In a multilocus sequence analysis, the similarity of VGXO14T and Vi1 to the closest type strain of the group, Pseudomonas pachastrellae, was 90.4%, which was lower than the threshold of 97% established to discriminate species in the Pseudomonas genus. The DNA–DNA hybridisation similarity between strains VGXO14T and Vi1 was 79.6%, but below 70% with the type strains in the P. pertucinogena group. Therefore, the strains should be classified within the genus Pseudomonas as a novel species, for which the name Pseudomonas aestusnigri is proposed. The type strain is VGXO14T (=CCUG 64165T = CECT 8317T).  相似文献   

4.
Two Gram-staining-negative, moderately halophilic bacteria, strains M1-18T and L1-16, were isolated from a saltern located in Huelva (Spain). They were motile, strictly aerobic rods, growing in the presence of 3–25% (w/v) NaCl (optimal growth at 7.5–10% [w/v] NaCl), between pH 4.0 and 9.0 (optimal at pH 6.0–7.0) and at temperatures between 15 and 40 °C (optimal at 37 °C). Phylogenetic analysis based on 16S rRNA gene sequence comparison showed that both strains showed the higher similarity values with Chromohalobacter israelensis ATCC 43985T (95.2–94.8%) and Chromohalobacter salexigens DSM 3043T (95.0–94.9%), and similarity values lower than 94.6% with other species of the genera Chromohalobacter, Kushneria, Cobetia or Halomonas. Multilocus sequence analysis (MLSA) based on the partial sequences of atpA, rpoD and secA housekeeping genes indicated that the new isolates formed an independent and monophyletic branch that was related to the peripheral genera of the family Halomonadaceae, Halotalea, Carnimonas and Zymobacter, supporting their placement as a new genus of the Halomonadaceae. The DNA–DNA hybridization between both strains was 82%, whereas the values between strain M1-18T and the most closely related species of Chromohalobacter and Kushneria were equal or lower to 48%. The major cellular fatty acids were C18:1ω7c/C18:1ω6c, C16:0, and C16:1ω7c/C16:1ω6c, a profile that differentiate this new taxon from species of the related genera. We propose the placement of both strains as a novel genus and species, within the family Halomonadaceae, with the name Larsenia salina gen. nov., sp. nov. The type strain is M1-18T (= CCM 8464 = CECT 8192T = IBRC-M 10767T = LMG 27461T).  相似文献   

5.
Herein I compare the relative importance of preference for structurally complex habitat against avoidance of competitors and predators in two benthic fishes common in the Gulf of Mexico. The code goby Gobiosoma robustum Ginsburg and clown goby Microgobius gulosus (Girard) are common, ecologically similar fishes found throughout the Gulf of Mexico and in the southeastern Atlantic Ocean. In Florida Bay, these fishes exhibit habitat partitioning: G. robustum is most abundant in seagrass-dominated areas while M. gulosus is most abundant in sparsely vegetated habitats. In a small-scale field survey, I documented the microhabitat use of these species where their distributions overlap. In a series of laboratory experiments, I presented each species with structured (artificial seagrass) versus nonstructured (bare sand) habitats and measured their frequency of choosing either habitat type. I then examined the use of structured versus nonstructured habitats when the two species were placed together in a mixed group. Finally, I placed a predator (Opsanus beta) in the experimental aquaria to determine how its presence influenced habitat selection. In the field, G. robustum was more abundant in seagrass and M. gulosus was more abundant in bare mud. In the laboratory, both species selected grass over sand in allopatry. However, in sympatry, M. gulosus occupied sand more often when paired with G. robustum than when alone. G. robustum appears to directly influence the habitat choice of M. gulosus: It seems that M. gulosus is pushed out of the structured habitat that is the preferred habitat of G. robustum. Thus, competition appears to modify the habitat selection of these species when they occur in sympatry. Additionally, the presence of the toadfish was a sufficient stimulus to provoke both M. gulosus and G. robustum to increase their selection for sand (compared to single-species treatments). Distribution patterns of M. gulosus and G. robustum likely result from a synthesis of various biotic and abiotic filters, including physiological tolerances to environmental factors, dispersal ability of larvae, and availability of food. Selection for structural complexity, competition, and presence of predators may further define the resulting pattern of distribution observed in the field.  相似文献   

6.
The composition of essential oils hydrodistilled from 19 samples of inflorescences and leaves of Achillea millefolium L. plants, which were transferred from 14 natural habitats in Lithuania to the field collection, is reported. Total content of oil was 0.15–0.55% in inflorescences and 0.06–0.19% (v/w) in leaves. In total 117 compounds were identified positively or tentatively. Data obtained clearly indicate the presence of a remarkable chemical polymorphism within the population of A. millefolium in Lithuania. The content of the major constituents in the oils from inflorescences varied in the following ranges: β-pinene, 0.33–62.29%; β-myrcene, 0.05–69.76%; α-phelandrene, 0.13–29.96%; 1,8-cineole, 2.30–21.57%; and chamazulene, 0.08–30.70%. According to the major components the essential oils' six chemotypes of A. millefolium were defined.  相似文献   

7.
The potential allelopathic effects of 14 stilbenoids and five flavonoids, isolated from leaves of Carex distachya Desf., were evaluated on the seed germination and seedling growth of three coexisting Mediterranean species (Dactylis hispanica, Petrorhagia velutina, and Phleum subulatum). The structures of the metabolites have been elucidated on the basis of their spectroscopic features (1D and 2D NMR experiments and EI–MS and ESI–MS data). The bioassays showed species-specific effects of the metabolites from C. distachya, specially on the plant growth (root and shoot elongation) which resulted significantly stimulated or inhibited at 10−4 M concentration. The effects on root elongation is generally greater than the shoot growth at all the tested concentrations (10−4–10−8 M). Cluster of biological data showed interesting relationships between the chemical structures of the compounds and their biological effects.  相似文献   

8.
Phenolic compounds were studied in the culms of five bamboo species collected in China: Yushania chungii, Fargesia robusta, Fargesia denudata, Fargesia rufa and Fargesia scabrida. All the species are eaten by giant panda (Ailuropoda melanoleuca). The culms contained phenolic acids and flavonoids in small concentrations, except for F. robusta, which did not contain flavonoids in detectable amounts. The species differed from each other in their phenolic composition. For example, F. rufa with the highest number of compounds clearly differed from other species. There were also differences among sampling sites, which reflect the differences among genotypes. Furthermore, there were clear ontogenetic differences in the culms: some compounds were present in mature culms but not in young (1–2 year old) culms, while the concentrations of other compounds decreased with increasing age. Over all, the composition and concentrations of soluble phenolic compounds in the bamboo culms were affected by species, age and site.  相似文献   

9.
Differentiation of freshly separated (uncultured) and cultured cyanobionts of Azolla species was carried out employing morphological markers and profiles generated using SDS-PAGE and PCR–RFLP of 16S rRNA. The cell dimensions of vegetative cells, heterocysts and heterocyst frequency (25–28%) of the freshly separated cyanobionts were distinctly higher than that those recorded for the cultured cyanobionts (7–10%). The SDS-PAGE profiles of whole cell proteins of cultured cyanobionts (comprising 28–30 bands) and those of freshly separated cyanobionts (consisting of 6–10 bands) exhibited distinct differences and unique bands. AluI was able to discriminate freshly separated cyanobionts from cultured cyanobionts of same species of Azolla. The profiles of cyanobionts (freshly separated and cultured) of A. rubra (RU6503) generated using the restriction enzyme AluI were distinctly different from other cyanobionts and unique bands were observed in both cyanobionts. The cultured cyanobionts from A. microphylla (MI4018), A. filiculoides (FI1001) and A. pinnata (PP7001) showed the presence of a distinct band of 450, 622 and 307 bp, respectively. Three common bands of 500, 400 and 275 bp were recorded in the AluI restriction profiles of all the freshly separated cyanobionts. 16S rRNA PCR–RFLP analyses confirmed the existence of primary and secondary cyanobionts in the leaf cavities of different Azolla species. These techniques can be utilized for discriminating between freshly separated and cultured cyanobionts of Azolla and provide reliable fingerprints for these strains.  相似文献   

10.
Fatty acid composition of nine species of Salvia, naturally growing in Turkey was determined: Salvia syriaca, Salvia potentillifolia, Salvia candidissima ssp. occidentalis, Salvia macrochlamys, Salvia poculata, Salvia tomentosa, Salvia recognita, Salvia virgata and Salvia ceratophylla. The main compounds were found to be linoleic acid (18:2; 24.3–69.2%), linolenic acid (18:3; 0.6–40.8%), oleic acid (18:1; 8.3–31.0%), palmitic acid (16:0; 3.8–21.0%) and stearic acid (18:0; 1.8–5.2%). Fatty acid composition of Salvia seed oils could be used as a chemotaxonomical marker.  相似文献   

11.
Four new polyhydroxylated sterols, named halicrasterols A–D (14), together with six known analogs (510) were isolated from the marine sponge Haliclona crassiloba. Compounds 1 and 2 represented rare examples of steroids featuring 17(20)E-double bonds. The structures of 110 were elucidated by spectroscopic analysis and comparison with reported data. This is the first report of a steroid profile for this species. The antimicrobial activities of 110 were evaluated against a panel of bacterial and fungal strains in vitro, and compounds 4 and 9 showed moderate activity against some of the Gram-positive strains with MICs ranging from 4 to 32 μg/mL.  相似文献   

12.
Nine (1–9) and seven (1–6, 10) compounds were isolated from the fruits of Sonneratia caseolaris and Sonneratia ovata, respectively. Their structures were identified by comparing their MS and NMR data as well as the physical properties with the literature. All the isolated compounds were screened against a rat glioma C-6 cell line using the MTT assay method; only compounds (-)-(R)-nyasol (1), (-)-(R)-4′-O-methylnyasol (2) and maslinic acid (6) were found to show moderate cytotoxic activity. Our findings from these two kinds of fruits can be used as a foundation for further chemotaxonomic studies on Sonneratia species. The nor-lignans (1, 2) and 6H-benzo[b,d]pyran-6-one derivatives (3, 4) were isolated from this genus for the first time, indicating that these two classes of compounds may tentatively be considered as taxonomic markers for Sonneratia genus.  相似文献   

13.
Galα1–4Gal is typically found in mammalian glycolipids in small quantities, and recognized by some pathogens, such as uropathogenic Escherichia coli. In contrast, glycoproteins containing Galα1–4Gal were rarely found in vertebrates except in a few species of birds and amphibians until recently. However, we had previously reported that pigeon (Columba livia) egg white and serum glycoproteins are rich in N-glycans with Galα1–4Gal at non-reducing termini. Our investigation with egg white glycoproteins from 181 avian species also revealed that the distribution of (Galα1–4Gal)-containing glycoproteins was not rare among avians, and is correlated with the phylogeny of birds. The differentiated expression was most likely emerged at earlier stage of diversification of modern birds, but some birds might have lost the facility for the expression relatively recently.  相似文献   

14.
In a chemosystematic investigation of the fruits of Daucus carota L. we have isolated eleven compounds including four guaiane type sesquiterpenes (14), one eudesmane type sesquiterpene (5) and six other constituents (611). Among them, the five compounds 15 all are new for Daucus carota L. and all have been reported in Torilis japonica. The results show that Daucus carota L. is chemically similar with Torilis japonica providing chemical evidence for both species used as the same traditional Chinese medicine “nanheshi”.  相似文献   

15.
Two new bufadienolide glycosides with an A/B trans ring structure, 14β,16β-dihydroxy-3β-(β-d-glucopyranosyloxy)-5α-bufa-20,22-dienolide (1), and 14β,16β-dihydroxy-3β-[β-d-glucopyranosyl-(1→4)-(β-d-glucopyranosyloxy)]-5α-bufa-20,22-dienolide (2), two known ecdysteroids (polypodine B and 20-hydroxyecdysone) (3-4), and six known bufadienolide and its glycosides with 5β-OH (hellebrigenin, 16β-hydroxyhellebrigenin-3-O-α-l-rhamnoside, hellebrigenin 3-O-β-d-glucoside, hellebrin, 16β-hydroxyhellebrigenin-3-O-β-d-glucoside, and deglucohellebrin) (5-10) were isolated from the rhizomes of Helleborus thibetanus. The structures of compounds 1 and 2 were elucidated using various spectroscopic methods. All compounds were reported for the first time from the title plant and their chemotaxonomic significance for the genus Helleborus was discussed.  相似文献   

16.
Eight compounds (1-7b) were isolated from the aerial parts of S. halepense in present investigation and five of them (2, 5, 6, 7a, 7b) were firstly reported from this species. The two rare diastereomeric flavonolignans tricin-4'-O-(threo-β-guaiacylglyceryl) ether and tricin-4'-O-(erythro-β-guaiacylglyceryl) ether is from Sorghum genus for the first time. The chemotaxonomic significance of these compounds was summarized.  相似文献   

17.
The qualitative and quantitative composition of the essential oils obtained from the inflorescences of Achyrocline flaccida (Asteraceae) has been investigated for the first time. Plant material was collected from eleven locations in Argentina. The essential oils were obtained by hydrodistillation (0.1–0.8% v/w, dried material) and analyzed by GC–FID–MS. Eighty-three compounds were identified representing more than the 90% of the oils. The major components were α-pinene and β-caryophyllene. Statistical analysis was performed in order to evaluate the variability of the essential oils analyzed. Two groups were formed reflecting only quantitative differences in the content of major compounds. The chemical pattern of essential oils observed for A. flaccida is similar to other Achyrocline species studied, except Achyrocline hyperchlora.  相似文献   

18.
The phylogeny and taxonomic position of slow-growing Genista tinctoria rhizobia from Poland, Ukraine and England were estimated by comparative 16S rDNA, atpD, and dnaK sequence analyses, PCR-RFLP of 16S rDNA, DNA G + C content, and DNA–DNA hybridization. Each core gene studied placed the G. tinctoria rhizobia in the genus Bradyrhizobium cluster with unequivocal bootstrap support. G. tinctoria symbionts and bradyrhizobial strains shared 96–99% similarity in 16S rDNA sequences. Their similarity for atpD and dnaK sequences was 93–99% and 89–99%, respectively. These data clearly showed that G. tinctoria rhizobia belonged to the genus Bradyrhizobium. 16S rDNA sequence analysis was in good agreement with the results of the PCR-RFLP of the 16S rRNA gene. Although the tested strains formed separate lineages to the reference bradyrhizobia their RFLP 16S rDNA patterns were quite similar. The genomic DNA G + C content of three G. tinctoria rhizobia was in the range from 60.64 to 62.83 mol%. Data for species identification were obtained from DNA–DNA hybridization experiments. G. tinctoria microsymbionts from Poland were classified within Bradyrhizobium japonicum genomospecies based on 56–82% DNA–DNA similarity.  相似文献   

19.
Chemical and molecular analyses were performed for the characterization of Juniperus brevifolia (Seub.) Antoine (Cupressaceae), an Azorean endemic species. Forty-two individual twig samples were collected at seven Azorean islands (S. Miguel, Terceira, S. Jorge, Pico, Faial, Flores and Corvo). Volatile compounds were isolated by distillation–extraction and analyzed by GC and GC-MS. The volatile oils consisted mainly of monoterpene hydrocarbons (84–96%), limonene (33–87%), α-pinene (1–48%) and β-myrcene (1–5%) being the main components, and cluster analysis showed a high correlation among all samples (Scorr = 0.84). DNA fingerprinting was performed using thirty-seven RAPD and eleven ISSR primers, generating 881 polymorphic bands. Cluster analysis showed a moderate correlation among accessions (SDice = 0.43), which grouped according to their geographical origin. Chemical and molecular data did not show superimposable clustering profiles. Although the molecular approaches tested were useful in assessing genetic diversity, no straight correlation could be drawn between chemical and molecular data sets.  相似文献   

20.
The family Trigonalyidae is considered to be one of the most basal lineages in the suborder Apocrita of Hymenoptera. Here, we determine the first complete mitochondrial genome of the Trigonalyidae, from the species Taeniogonalos taihorina (Bischoff, 1914). This mitochondrial genome is 15,927 bp long, with a high A + T-content of 84.60%. It contains all of the 37 typical animal mitochondrial genes and an A + T-rich region. The orders and directions of all genes are different from those of previously reported hymenopteran mitochondrial genomes. Eight tRNA genes, three protein-coding genes and the A + T-rich region were rearranged, with the dominant gene rearrangement events being translocation and local inversion. The arrangements of three tRNA clusters, trnYtrnMtrnItrnQ, trnWtrnL2trnC, and trnHtrnAtrnRtrnNtrnStrnEtrnF, and the position of the cox1 gene, are novel to the Hymenoptera, even the insects. Six long intergenic spacers are present in the genome. The secondary structures of the RNA genes are normal, except for trnS2, in which the D-stem pairing is absent.  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