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.
The presence of repeated DNA (short tandemly repeated repetitive [STRR] and long tandemly repeated repetitive [LTRR]) sequences in the genome of cyanobacteria was used to generate a fingerprint method for symbiotic and free-living isolates. Primers corresponding to the STRR and LTRR sequences were used in the PCR, resulting in a method which generate specific fingerprints for individual isolates. The method was useful both with purified DNA and with intact cyanobacterial filaments or cells as templates for the PCR. Twenty-three Nostoc isolates from a total of 35 were symbiotic isolates from the angiosperm Gunnera species, including isolates from the same Gunnera species as well as from different species. The results show a genetic similarity among isolates from different Gunnera species as well as a genetic heterogeneity among isolates from the same Gunnera species. Isolates which have been postulated to be closely related or identical revealed similar results by the PCR method, indicating that the technique is useful for clustering of even closely related strains. The method was applied to nonheterocystus cyanobacteria from which a fingerprint pattern was obtained.  相似文献   

2.
Early events during the establishment of the Gunnera/Nostoc symbiosis   总被引:1,自引:0,他引:1  
The symbiosis between Gunnera and Nostoc was reconstituted using G. chilensis Lam. and G. manicata Linden, respectively, and three different Nostoc strains. Six stages characterised by specific modifications in both the cyanobiont and the host were recognised during the infection process. Mucilage-secreting stem glands developed on the Gunnera stems independent of the presence of cyanobacteria (Stage I). Soon after addition of the Nostoc isolates to the plant apices, an abundant differentiation of motile hormogonia commenced. The cyanobacteria accumulated in the mucilage on the surface of the gland (Stage II), and the hormogonia then proceeded into the stem tissue through intercellular channels (Stage III). At the channel bases, Nostoc was detected between the cell walls of small, densely cytoplasmic Gunnera cells and also in elaborate folds of these (Stage IV). The Gunnera cell walls subsequently dissolved adjacent to the cyanobacteria and Nostoc entered the host cells (Stage V). Once the intracellular association was formed, a high proportion of the vegetative Nostoc cells differentiated into heterocysts (Stage VI). Nostoc changed from being rich in inclusions (particularly cyanophycin) while on the gland surface into a comparatively non-storing form during penetration and the early intracellular stages. Bacteria were numerous on the gland surface, fewer in the channels, and were never detected within the Gunnera cells, indicating the existence of specific recognition mechanisms discriminating between conceivable microsymbionts. Mechanisms behind mutual adaptations and interactions between the two symbionts are discussed.The technical assistance of Anette Axen and Gary Wife is gratefully acknowledged. Financial support was provided by the Swedish Natural Science Research Council and the Hierta-Retzius foundations.  相似文献   

3.
In symbiotic systems, patterns of symbiont diversity and selectivity are crucial for the understanding of fundamental ecological processes such as dispersal and establishment. The lichen genus Nephroma (Peltigerales, Ascomycota) has a nearly cosmopolitan distribution and is thus an attractive model for the study of symbiotic interactions over a wide range of spatial scales. In this study, we analyze the genetic diversity of Nephroma mycobionts and their associated Nostoc photobionts within a global framework. The study is based on Internal Transcribed Spacer (ITS) sequences of fungal symbionts and tRNALeu (UAA) intron sequences of cyanobacterial symbionts. The full data set includes 271 Nephroma and 358 Nostoc sequences, with over 150 sequence pairs known to originate from the same lichen thalli. Our results show that all bipartite Nephroma species associate with one group of Nostoc different from Nostoc typically found in tripartite Nephroma species. This conserved association appears to have been inherited from the common ancestor of all extant species. While specific associations between some symbiont genotypes can be observed over vast distances, both symbionts tend to show genetic differentiation over wide geographic scales. Most bipartite Nephroma species share their Nostoc symbionts with one or more other fungal taxa, and no fungal species associates solely with a single Nostoc genotype, supporting the concept of functional lichen guilds. Symbiont selectivity patterns within these lichens are best described as a geographic mosaic, with higher selectivity locally than globally. This may reflect specific habitat preferences of particular symbiont combinations, but also the influence of founder effects.  相似文献   

4.
5.
Element analysis using electron spectroscopic imaging (ESI) and electron energy loss spectroscopy (EELS) was performed in a symbiotic Nostoc sp. strain found in the upper stem tissue of Gunnera manicata, and in Nostoc PCC 9229, a free-living heterocyst-forming cyanobacterium able to enter into symbiosis with the angiosperm Gunnera in reconstitution experiments. ESI and EELS unequivocally identified the four elements nitrogen (N), sulphur (S), phosphorus (P) and oxygen (O) in different inclusion bodies of these biological specimens. High amounts of nitrogen were solely detected in huge cyanophycin granules in vegetative cells of the symbiotic Nostoc strain, whereas large polyphosphate bodies, containing high amounts of phosphorus, sulphur and oxygen, could be seen in the free-living Nostoc PCC 9229. The latter were usually not present or, when found, very small in vegetative cells of the cyanobiont.  相似文献   

6.
The diversity of the endophytic Nostoc symbionts of two thalloid bryophytes, the hornwort Anthoceros fusiformis and the liverwort Blasia pusilla, was examined using the tRNALeu (UAA) intron sequence as a marker. The results confirmed that many different Nostoc strains are involved in both associations under natural conditions in the field. The level of Nostoc diversity within individual bryophyte thalli varied, but single DNA fragments were consistently amplified from individual symbiotic colonies. Some Nostoc strains were widespread and were detected from thalli collected from different field sites and different years. These findings indicate a moderate level of spatial and temporal continuity in bryophyte-Nostoc symbioses.  相似文献   

7.
Gunnera plants have the unique ability to form endosymbioses with N2-fixing cyanobacteria, primarily Nostoc. Cyanobacteria enter Gunnera through transiently active mucilage-secreting glands on stems. We took advantage of the nitrogen (N)-limitation-induced gland development in Gunnera manicata to identify factors that may enable plant tissue to attract and maintain cyanobacteria colonies. Cortical cells in stems of N-stressed Gunnera plants were found to accumulate a copious amount of starch, while starch in the neighboring mature glands was nearly undetectable. Instead, mature glands accumulated millimolar concentrations of glucose (Glc) and fructose (Fru). Successful colonization by Nostoc drastically reduced sugar accumulation in the surrounding tissue. Consistent with the abundance of Glc and Fru in the gland prior to Nostoc colonization, genes encoding key enzymes for sucrose and starch hydrolysis (e.g. cell wall invertase, α-amylase, and starch phosphorylase) were expressed at higher levels in stem segments with glands than those without. In contrast, soluble sugars were barely detectable in mucilage freshly secreted from glands. Different sugars affected Nostoc’s ability to differentiate motile hormogonia in a manner consistent with their locations. Galactose and arabinose, the predominant constituents of polysaccharides in the mucilage, had little or no inhibitory effect on hormogonia differentiation. On the other hand, soluble sugars that accumulated in gland tissue, namely sucrose, Glc, and Fru, inhibited hormogonia differentiation and enhanced vegetative growth. Results from this study suggest that, in an N-limited environment, mature Gunnera stem glands may employ different soluble sugars to attract Nostoc and, once the cyanobacteria are internalized, to maintain them in the N2-fixing vegetative state.Nitrogen (N) is an essential element for plant growth, but availability of reduced N in the soil is often limiting. Representatives from a wide range of land plants have evolved the ability to form associations with N2-fixing microbes (Franche et al., 2009). Associations between rhizobia and legume plants are well-characterized examples of plant-bacterial N2-fixing symbioses. Unlike rhizobia, which generally exhibit narrow host ranges (Kistner and Parniske, 2002), N2-fixing cyanobacteria are able to form productive associations with a broad range of plants, including bryophytes (hornworts and liverworts), ferns (Azolla), gymnosperms (cycads), and angiosperms (Gunnera; for review, see Rai et al., 2000; Adams et al., 2006). Free-living cyanobacteria within the genus Nostoc can fix N in specialized microoxic cells called heterocysts. The ability of Nostoc to fix N independent of a host environment may facilitate the formation of symbioses with a wide range of plants. Understanding the physiological conditions that enable a plant host to enter into symbiotic associations with cyanobacteria may allow us to extend the benefit of biological N fixation to crops outside the legume family.Nostoc has the ability to differentiate not only into filaments bearing heterocysts but also into transiently motile filaments, known as hormogonia, which enable the cyanobacteria to enter plants (Meeks and Elhai, 2002). Nostoc can be induced to form hormogonia by different environmental stimuli and by a hormogonia-inducing factor released from N-stressed host plants (Meeks and Elhai, 2002; Adams et al., 2006). The attraction of hormogonia to plants is much less specific than that of rhizobia. Hormogonia are attracted to root extracts from either host or nonhost plants and even to certain simple sugars, such as Ara, Glc, and Gal (Nilsson et al., 2006). After entering a plant host, hormogonia revert back to filaments with N2-fixing heterocysts. Inside the host, further hormogonia formation is suppressed, and heterocysts appear at a frequency of about 30% to 40%, 3- to 4-times higher than that found in free-living Nostoc (Meeks and Elhai, 2002). Although free-living Nostoc species can support N2 fixation through photosynthesis, under symbiotic conditions they rely on photosynthate from the host plant. In general, the sugars (Suc, Glc, and Fru) known to support heterotrophic growth in the dark by free-living cyanobacteria coincide with those that support nitrogenase activity in Nostoc-plant associations (Meeks and Elhai, 2002). However, the Nostoc-Gunnera association may be exceptional; only Glc and Fru have been shown to sustain nitrogenase activities (Man and Silvester, 1994; Wouters et al., 2000), although Suc anddextrin were able to keep Nostoc alive without light (Wouters et al., 2000). It is evident from cyanobacterial studies that the plant hosts have evolved the ability to regulate cyanobacterial growth and differentiation during symbiotic associations (Meeks and Elhai, 2002).However, because most studies of plant-cyanobacterial associations have focused on the cyanobacterial partner (e.g. Wang et al., 2004; Ekman et al., 2006), the mechanisms through which plant hosts attract, internalize, and maintain cyanobacteria remain to be elucidated (Adams et al., 2006).The Nostoc-Gunnera association is an ideal system with which to study plant-cyanobacteria symbioses, not only because Gunnera is the only genus of angiosperms known to form endosymbioses with N2-fixing cyanobacteria but also because the association between the two can be readily established in the laboratory (Bergman et al., 1992; Chiu et al., 2005). Nostoc hormogonia enter Gunnera plants through specialized glands located on the stem. As the gland matures, it secretes polysaccharide-rich mucilage that attracts cyanobacteria (Nilsson et al., 2006), supports their growth on the gland surface (Towata, 1985; Chiu et al., 2005), and permits further hormogonia differentiation (Rasmussen et al., 1994). From there, hormogonia enter the gland and penetrate cells near the base of the gland in the stem (Bonnett, 1990; Bergman et al., 1992). Although each gland is only transiently capable of accepting cyanobacteria, new glands continue to form on the stem at the base of each new leaf.In contrast to the development of nodules in legumes, which requires a complex exchange of signals between the two symbiotic partners (Cooper, 2007), stem gland development in Gunnera takes place in the absence of cyanobacteria (Bonnett, 1990). N limitation, however, is a prerequisite for stem gland development (Chiu et al., 2005), as it is for nodulation (Barbulova et al., 2007). We have taken advantage of the N-deficiency-induced gland development in G. manicata to identify factors that enable Gunnera to form endosymbiosis with cyanobacteria. This study investigated changes in the carbohydrate metabolism during Gunnera gland development and discovered that tissue in the mature glands accumulated high levels of soluble sugars prior to the arrival of cyanobacteria. In agreement with this finding, several key genes encoding enzymes for starch/Suc hydrolysis were expressed at higher levels in the gland compared to the stem. Furthermore, we found that various sugars cyanobacteria may encounter as they approach Gunnera glands as opposed to those they would encounter within plant cells differentially affected Nostoc’s ability to form motile hormogonia.  相似文献   

8.
This study evaluated the feasibility of using the Ribulose Bisphosphate Carboxylase Large subunit gene (rbcL) and the Internal Transcribed Spacers 1 and 2 of the nuclear rDNA (nuITS1 and nuITS2) markers for identifying a very diverse, albeit poorly known group, of green microalgae from neotropical inland waters. Fifty-one freshwater green microalgae strains isolated from Brazil, the largest biodiversity reservoir in the neotropics, were submitted to DNA barcoding. Currently available universal primers for ITS1-5.8S-ITS2 region amplification were sufficient to successfully amplify and sequence 47 (92%) of the samples. On the other hand, new sets of primers had to be designed for rbcL, which allowed 96% of the samples to be sequenced. Thirty-five percent of the strains could be unambiguously identified to the species level based either on nuITS1 or nuITS2 sequences’ using barcode gap calculations. nuITS2 Compensatory Base Change (CBC) and ITS1-5.8S-ITS2 region phylogenetic analysis, together with morphological inspection, confirmed the identification accuracy. In contrast, only 6% of the strains could be assigned to the correct species based solely on rbcL sequences. In conclusion, the data presented here indicates that either nuITS1 or nuITS2 are useful markers for DNA barcoding of freshwater green microalgae, with advantage for nuITS2 due to the larger availability of analytical tools and reference barcodes deposited at databases for this marker.  相似文献   

9.
A phylogenetic analysis of selected symbiotic Nostoc strain sequences and available database 16S rDNA sequences of both symbiotic and free-living cyanobacteria was carried out using maximum likelihood and Bayesian inference techniques. Most of the symbiotic strains fell into well separated clades. One clade consisted of a mixture of symbiotic and free-living isolates. This clade includes Nostoc sp. strain PCC 73102, the reference strain proposed for Nostoc punctiforme. A separate symbiotic clade with isolates exclusively from Gunnera species was also obtained, suggesting that not all symbiotic Nostoc species can be assigned to N. punctiforme. Moreover, isolates from Azolla filiculoides and one from Gunnera dentata were well nested within a clade comprising most of the Anabaena sequences. This result supports the affiliation of the Azolla isolates with the genus Anabaena and shows that strains within this genus can form symbioses with additional hosts. Furthermore, these symbiotic strains produced hormogonia, thereby verifying that hormogonia formation is not absent in Anabaena and cannot be used as a criterion to distinguish it from Nostoc.The GenBank accession numbers for the cyanobacterial 16S rRNA gene sequences determined in this study are AY742447-AY742454.  相似文献   

10.

Background

Although Cryptococcus laurentii has been considered saprophytic and its taxonomy is still being described, several cases of human infections have already reported. This study aimed to evaluate molecular aspects of C. laurentii isolates from Brazil, Botswana, Canada, and the United States.

Methods

In this study, 100 phenotypically identified C. laurentii isolates were evaluated by sequencing the 18S nuclear ribosomal small subunit rRNA gene (18S-SSU), D1/D2 region of 28S nuclear ribosomal large subunit rRNA gene (28S-LSU), and the internal transcribed spacer (ITS) of the ribosomal region.

Results

BLAST searches using 550-bp, 650-bp, and 550-bp sequenced amplicons obtained from the 18S-SSU, 28S-LSU, and the ITS region led to the identification of 75 C. laurentii strains that shared 99–100% identity with C. laurentii CBS 139. A total of nine isolates shared 99% identity with both Bullera sp. VY-68 and C. laurentii RY1. One isolate shared 99% identity with Cryptococcus rajasthanensis CBS 10406, and eight isolates shared 100% identity with Cryptococcus sp. APSS 862 according to the 28S-LSU and ITS regions and designated as Cryptococcus aspenensis sp. nov. (CBS 13867). While 16 isolates shared 99% identity with Cryptococcus flavescens CBS 942 according to the 18S-SSU sequence, only six were confirmed using the 28S-LSU and ITS region sequences. The remaining 10 shared 99% identity with Cryptococcus terrestris CBS 10810, which was recently described in Brazil. Through concatenated sequence analyses, seven sequence types in C. laurentii, three in C. flavescens, one in C. terrestris, and one in the C. aspenensis sp. nov. were identified.

Conclusions

Sequencing permitted the characterization of 75% of the environmental C. laurentii isolates from different geographical areas and the identification of seven haplotypes of this species. Among sequenced regions, the increased variability of the ITS region in comparison to the 18S-SSU and 28S-LSU regions reinforces its applicability as a DNA barcode.  相似文献   

11.
The phycobiliprotein phycoerythrin was localized in symbiotic and free-living Nostoc of the cycad Cycas using immunocytochemistry. In symbiotic Nostoc, phycoerythrin was associated with the thylakoid membranes of vegetative cells and absent from heterocysts. Similar cellular/subcellular localization was observed between symbiotic Nostoc and the free-living Cycas isolate Nostoc 7422.  相似文献   

12.
The partners of the symbiotic association between Anthoceros punctatus L. and Nostoc spp. have been cultured separately in a pure state. The symbiotic association was reconstituted following dual culture in liquid Anthoceros growth medium with a variety of axenic Nostoc isolates and mutant strains. The heterocyst frequency of competent Nostoc strains increased four- to fivefold when in symbiotic association relative to free-living N2-grown cultures. Dinitrogen fixation by symbiotic Nostoc supported the growth of Anthoceros tissue, although this growth was nitrogen-limited relative to that supported by exogenous ammonium. When the association was reconstituted in the presence of two or three wild-type and mutant Nostoc strains some of these strains were found to compete in infection of Anthoceros tissue and a fraction of the symbiotic Nostoc colonies contained more than one strain. Exogenous ammonium did not affect infection, but repressed development of the symbiotic Nostoc colonies in Anthoceros tissue, and symbiotic Nostoc in N2-grown Anthoceros tissue appeared to regress from the symbiotic state in the presence of exogenous ammonium. The results show that the Anthoceros-Nostoc symbiotic association is amenable to specific experimental manipulations; their implications are discussed with respect to infection of Anthoceros tissue and control of the development of symbiotic Nostoc.  相似文献   

13.
Scirtothrips dorsalis Hood is a cosmopolitan and polyphagous thrips species. Recently, a novel strain of S. dorsalis attacking capsicum crops was found in Japan. A molecular phylogenetic analysis using mitochondrial cytochrome c oxidase subunit I sequences revealed that the capsicum-associated populations were genetically different from Japanese native strains and were closely related to Southeast Asian populations. We named the capsicum-associated populations “strain C” and the Japanese native ones “strain YT”. A total of 10 haplotypes were found in strain C and 26 in strain YT. To differentiate the two strains, we developed a multiplex-PCR method using the ribosomal ITS2 region.  相似文献   

14.
Specimens sampled in Central Mongolia have been examined for the intraspecific polymorphism of the nucleotide sequences of ribosomal DNA internal transcribed spacers (ITS1 + 5.8S + ITS2) of the parasitic nematode Haemonchus contortus. A considerable diversity of haplotypes differing in the nucleotide composition of this DNA region has been observed. The phylogenetic relationships between the haplotypes detected in Central Mongolia and the corresponding sequences from other parts of the nematode distribution area (deposited with the NCBI GenBank) have been analyzed. Significantly different sequences have been found along with the haplotypes already observed in H. contortus or differing from them by one-two nucleotides.  相似文献   

15.
Most molecular ecological studies of arbuscular mycorrhizal fungi (AMF) have been based on the rRNA gene sequences. However, information about intraspecific nucleotide variation is still limited in these fungi. In this study, we calculated the inter- and intrasporal nucleotide variation of Diversispora sp. EE1 using 78 cloned sequences from four spores within a ca 4960 bp fragment of the nuclear ribosomal operon spanning the near full length small ribosomal subunit (SSU) rRNA gene, the full internal transcribed spacer (ITS: ITS1-5.8S-ITS2) and ca 2740 bp of the large ribosomal subunit (LSU) rRNA gene. Data for each marker region (SSU, ITS and LSU) originated from the very same spores. Sequence variation resulting from point mutations and small indels was recorded in all regions. Highest sequence variation was observed in the ITS region at both the inter- and intrasporal levels. The ITS1 component was more variable than ITS2, whilst the 5.8S gene was the least variable component of the ITS region. Evolutionary divergence of gene copies between spores was intermediate for the LSU and lowest for the SSU. The SSU and the LSU genes had relatively similar evolutionary divergence per spore. Sequence variant richness was not exhaustive for any of the marker regions, indicating that multiple sequences per spore from multiple spores are needed when characterizing a species. This study provides reference sequences for ecological studies, permitting identification of AMF using any of the ribosomal regions or primer systems.  相似文献   

16.
Omphalina basidiolichens are obligate mutualistic associations of a fungus of the genus Omphalina (the exhabitant) and a unicellular green alga of the genus Coccomyxa (the inhabitant). It has been suggested that symbiotic inhabitants have a lower rate of genetic change compared to exhabitants because the latter are more exposed to abiotic environmental variation and competition from other organisms. In order to test this hypothesis we compared substitution rates in the nuclear ribosomal internal transcribed spacer region (ITS1, 5.8S, ITS2) among fungal species with rates among their respective algal symbionts. To ensure valid comparisons, only taxon pairs (12) with a common evolutionary history were used. On average, substitution rates in the ITS1 portion of Omphalina pairs were 27.5 times higher than rates in the corresponding pairs of Coccomyxa since divergence from their respective ancestor at the base of the Omphalina/Coccomyxa lineage. Substitution rates in the 5.8S and the ITS2 portions were 2.4 and 18.0 times higher, respectively. The highest rate difference (43.0) was found in the ITS1 region. These are, to our knowledge, the highest differences of substitution rates reported for symbiotic organisms. We conclude that the Omphalina model system conforms to the proposed hypothesis of lower substitution rates in the inhabitant, but that the mode of transmission of the inhabitant (vertical versus horizontal) could be a prevailing factor in the regulation of unequal rates of nucleotide substitution between co-evolving symbionts. Our phylogenetic study of Coccomyxa revealed three main lineages within this genus, corresponding to free-living Coccomyxa, individuals isolated from basidiolichens Omphalina and Coccomyxa isolated from ascolichens belonging to the Peltigerales.  相似文献   

17.
We previously reported the occurrence of genetically‐diverse symbiotic dinoflagellates (zooxanthellae) within and between 7 giant clam species (Tridacnidae) from the Philippines based on the algal isolates' allozyme and random amplified polymorphic DNA (RAPD) patterns. We also reported that these isolates all belong to clade A of the Symbiodinium phylogeny with identical 18S rDNA sequences. Here we extend the genetic characterization of Symbiodinium isolates from giant clams and propose that they are conspecific. We used the combined DNA sequences of the internal transcribed spacer (ITS)1, 5.8S rDNA, and ITS2 regions (rDNA‐ITS region) because the ITS1 and ITS2 regions evolve faster than 18S rDNA and have been shown to be useful in distinguishing strains of other dinoflagellates. DGGE of the most variable segment of the rDNA‐ITS region, ITS1, from clonal representatives of clades A, B, and C showed minimal intragenomic variation. The rDNA‐ITS region shows similar phylogenetic relationships between Symbiodinium isolates from symbiotic bivalves and some cnidarians as does 18S rDNA, and that there are not many different clade A species or strains among cultured zooxanthellae (CZ) from giant clams. The CZ from giant clams had virtually identical sequences, with only a single nucleotide difference in the ITS2 region separating two groups of isolates. These data suggest that there is one CZ species and perhaps two CZ strains, each CZ strain containing individuals that have diverse allozyme and RAPD genotypes. The CZ isolated from giant clams from different areas in the Philippines (21 isolates, 7 clam species), the Australian Great Barrier Reef (1 isolate, 1 clam species), Palau (8 isolates, 7 clam species), and Okinawa, Japan (1 isolate, 1 clam species) shared the same rDNA‐ITS sequences. Furthermore, analysis of fresh isolates from giant clams collected from these geographical areas shows that these bivalves also host indistinguishable clade C symbionts. These data demonstrate that conspecific Symbiodinium genotypes, particularly clade A symbionts, are distributed in giant clams throughout the Indo‐Pacific.  相似文献   

18.
Endosymbiotic green algae of Japanese Paramecium bursaria were phylogenetically analyzed based on DNA sequences from the ribosomal DNA operon (18S rDNA, ITS1, 5.8S rDNA, and ITS2). Phylogenetic trees constructed using 18S rDNA sequences showed that the symbionts belong to the Chlorella sensu stricto (Trebouxiophyceae) group. They are genetically closer to the C. vulgaris Beijerinck group than to C. kessleri Fott et Nováková as proposed previously. Branching order in C. vulgaris group was unresolved in 18S rDNA trees. Compared heterogeneities of 18S rDNA, ITS1, 5.8S r, and ITS2 among symbionts and two Chlorella species, indicated that the ITS2 region (and probably also ITS1) is better able to resolve phylogenetic problems in such closely related taxa. All six symbiotic sequences obtained here (approximately 4000-bp sequences of 18S rDNA, ITS1, 5.8S rDNA, and ITS2) were completely identical in each, strongly suggesting a common origin.  相似文献   

19.
Molecular analysis of green-tide-forming macroalgae in the Yellow Sea   总被引:2,自引:0,他引:2  
In the summer of 2008, free-floating green algae bloomed in the Yellow Sea. Samples were collected in a wide area (119°32′-122°00′E, 32°25′-36°49′N). We calculated the sequence divergences of nuclear ITS, chloroplast rbcL, and psbA data of free-floating samples collected from the Yellow Sea and Ulvaceae from Europe and Japan. In the ITS sequence, 19 out of the 21 Yellow Sea samples of 2008 were identical to those of a sample taken at Qingdao in 2007. A low divergence (0.2%) was found in remaining two samples. Similar evidence was shown by pairwise distances of rbcL and psbA gene sequence data, implying the uniformity of the Yellow Sea blooms in 2007 and 2008. The ITS sequence of the Yellow Sea samples differed 8.1-10.8% from free-floating Enteromorpha or Ulva reported worldwide. ITS-based molecular phylogenetic results and rbcL sequence data grouped the free-floating alga in the Yellow Sea into one clade with Enteromorpha procera, Enteromorpha linza and Enteromorpha prolifera. Furthermore, both morphological characteristics and ribotype network of the ITS sequences imply that the blooming algae in 2007 and 2008 were E. prolifera. The haplotypes of the Yellow Sea free-floating E. prolifera are closely related to those from the Japanese coast but less to European and American algae.  相似文献   

20.
Genetic similarity among cyanobacteria of a morphological subgroup ofNostoc was evaluated through a comparison of several specific genes and the extent of DNA methylation. Four of six cyanobacteria were originally cultured from facultative symbioses with higher plants (Gunnera andEncephalartos); these and one free-living isolate had been identified or reputed to beN. punctiforme. No consistent correlation to species or symbiotic history was found from DNA hybridizations to genes coding for phycocyanin (cpcAB), allophycocyanin (apcAB), gas vesicle protein (gvpA1), and dinitrogenase reductase (nifH). One gene (gvpC) was not present, andgvpA1 was a single-copy gene in all strains. The gas vesicle genes were concluded to be potentially useful for broadly characterizingNostoc or at least this subgroup. Incubations ofNostoc genomic DNA with 22 restriction endonucleases indicated a high degree of methylation and similarity of its methylated DNA to that of other heterocystous cyanobacteria. The genetic variation of theNostoc isolates was judged to reflect primarily different soil origins.  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