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.
Parasitic dinoflagellates of the genus Amoebophrya infect and kill bloom‐forming dinoflagellates, including the toxic species Karlodinium micrum. Unlike non‐toxic hosts, K. micrum is partially resistant to infection, a trait that may be related to toxin production. Here we tested the hypothesis that parasitism of K. micrum is inversely related to toxin concentration in the culture medium. Time‐course studies were conducted to determine the influence of extracted toxin and toxin carrier (methanol) on host growth, parasite prevalence, and parasite load. Results indicate that methanol concentrations below 0.1% have no effect on these variables. When methanol concentration was maintained below 0.1%, extracted toxin equivalent to 100 to 10,000 K. micrum per ml had no effect on host abundance. We are currently analyzing sample to assess the fate of Amoebophrya dinospores when exposed to K. micrum toxin. We will also consider the effect of intracellular host toxin on parasite success, by examining the fate of Amoebophrya dinospores when inoculated to K. micrum cultures that express different levels of toxin production. Understanding the effect of toxins on parasite success will contribute to our knowledge of host‐parasite biology and provide insight into the role of dinoflagellate toxins as a defense against parasitism.  相似文献   

2.
The endoparasitic dinoflagellate Amoebophrya infects a number of free‐living marine dinoflagellates, including harmful algal bloom species. The parasitoid eventually kills its host and has been proposed to be a significant loss factor for dinoflagellate blooms in restricted coastal waters. For several decades, the difficulties of culturing host‐parasitoid systems have been a great obstacle for further research on the biology of Amoebophrya. Here, we established an Akashiwo sanguineaAmoebophrya sp. coculture from Chinese coastal waters and studied the parasitoid's generation time, dinospore survival and infectivity, as well as its host specificity. The lifespan of Amoebophrya sp. ex. A. sanguinea was approximately 58 h. The infective dinospores can survive up to 78 h in ambient waters but gradually lose their infectivity. The parasitoid was unable to infect other dinoflagellate species, its infection rate reached as high as 91% when the ratio of dinospores to host cells was 20:1. The high infectivity of dinospores suggests that the Amoebophrya strain was capable of removing a considerable fraction of host biomass within a short period, but that it is probably unable to maintain high infection levels under nonbloom conditions of its host, due to limited survival and time constraints in encountering host cells.  相似文献   

3.
Members of the Amoebophrya ceratii complex are endoparasitic dinoflagellates that parasitize a number of their dinoflagellate relatives, including toxic and/or harmful algal bloom-forming species. Despite many studies on the occurrence, prevalence, biology and molecular phylogeny of Amoebophrya spp., little attention has been given to toxin dynamics of host population following parasitism. Using Amoebophrya sp. infecting the paralytic shellfish toxin (PSP)-producing dinoflagellate Alexandrium fundyense, we addressed the following questions: (1) does parasitism by Amoebophrya sp. alter toxin content and toxin profiles of the dinoflagellate A. fundyense over the infection cycle? and (2) do parasite dinospores produced at the end of the infection cycle retain host toxins and thus potentially act as a vector to convey PSP toxin through the marine microbial food-web? Toxin time-course experiments showed that the PSP toxin contents did not vary significantly over the infection cycle, but mean toxin content for infected cultures was significantly higher than that for uninfected cultures. Host toxins were not detected in the free-living, dinospore stage of the parasite. Therefore, our results indicate that Amoebophrya sp. does not function as a vector for transferring PSP toxins to higher trophic levels. Rather, Amoebophrya infections appear to play an important role in maintaining healthy ecosystems by transforming potent toxins-producing dinoflagellates into non-toxic dinospores, representing “edible food” for consumers of the marine microbial food-web during toxic algal bloom event.  相似文献   

4.
The endoparasitic dinoflagellate Amoebophrya ceratii (Koeppen) Cachon uses a number of its free‐living relatives as hosts and may represent a species complex composed of several host‐specific parasites. Two thecate host–parasite systems [Amoebophrya spp. ex Alexandrium affine (Inoue and Fukuyo) Balech and ex Gonyaulax polygramma Stein], were used to test the hypothesis that two strains of Amoebophrya have a high degree of host specificity. To test this hypothesis, a series of cross‐infection experiments were conducted, with 10 thecate and three athecate dinoflagellate species as potential hosts. Surprisingly, the two strains of Amoebophrya lacked host specificity and had wider host ranges than previously recognized. Among the host species tested, Amoebophrya sp. ex Alexandrium affine was capable of infecting only species of genus Alexandrium (Alexandrium affine, Alexandrium catenella, and Alexandrium tamarense), while the parasite from Gonyaulax polygramma infected species covering five genera (Alexandrium, Gonyaulax, Prorocentrum, Heterocapsa, and Scripsiella). In the context of previous reports, these results suggest that host specificity of Amoebophrya strains varies from extremely species‐specific to rather unspecific, with specificity being stronger for strains isolated from athecate hosts. Information on host specificity of Amoebophrya strains provided here will be helpful in assessing the possibility of using these parasites as biological control agents for harmful algal blooms, as well as in defining species of Amoebophrya in the future.  相似文献   

5.
Members of Amoebophrya ceratii complex are known to infect a number of free‐living dinoflagellates including harmful algal bloom species. In August and October 2012, Amoebophrya infections during two bloom events of the dinoflagellate Cochlodinium polykrikoides were observed along southern coastal waters of Korea. Microscopic observations and molecular data revealed that two different Amoebophrya parasites infected the same host species. In addition, while one developed in the host's nucleus, the other in the host's cytoplasm. Phylogenetic analyses showed that the two parasites were not nested in the previously recognized “Amoebophrya ceratii complex clade”, which contained sequences of parasites infecting numerous dinoflagellate species. Instead, they branched as sister taxa to the isolate (possibly Amoebophrya) from radiolarians Hexacontium gigantheum. Our result indicates that the two Amoebophrya parasites infecting C. polykrikoides may be different species from those inside the “complex.”  相似文献   

6.
We examined the sterol profile of Karlodinium veneficum (D. Ballant.) J. Larsen, Akashiwo sanguinea (Hiraska) Ge. Hansen et Moestrup, Alexandrium tamarense (M. Lebour) Balech, Alexandrium affine (H. Inoue et Fukuyo) Balech, Gonyaulax polygramma F. Stein, and Gymnodinium instriatum (Freud. et J. J. Lee) Coats, along with their Amoebophyra parasites. There were no consistent sterol profiles that characterized the genus Amoebophyra. Instead, in five out of six comparisons, the host and parasite sterol profiles where highly correlated. The one exception, Amoebophyra sp. ex Alex. tamarense, was least like its host in sterol profile and also possessed the widest host range for infection. There was little correlation between host and parasite in fatty acid profiles, with the parasite being deficient in fatty acids characteristic of the plastid [e.g., 18:5(n‐3) associated with galactolipids of the thylakoids, as previously published by Adolf et al. (2007)]. Those hosts and parasites with sterol profiles dominated by desmethyl sterols were most sensitive to karlotoxin toxicity. In the host‐parasite pairs most sensitive to karlotoxin addition, recovery of the intact karlotoxin molecule was poorest. Given the sensitivity to karlotoxin, some species of Amoebophyra may avoid infection of K. veneficum.  相似文献   

7.
Amoebophrya is a marine parasite recently found to infect and kill bloom-forming dinoflagellates in the California Current System (CCS). However, it is unknown whether parasitism by Amoebophrya can control dinoflagellate blooms in major eastern boundary upwelling systems, such as the CCS. We quantified the abundance of a common bloom-forming species Akashiwo sanguinea and prevalence of its parasite (i.e., % infected cells) in surface water samples collected weekly from August 2005 to December 2008 at the Santa Cruz Wharf (SCW), Monterey Bay, CA. Additionally, we measured physical and chemical properties at the SCW and examined regional patterns of wind forcing and sea surface temperature. Relative abundance of the net phytoplankton species was also analyzed to discern whether or not parasitism influences net phytoplankton community composition. Epidemic infection outbreaks (>20% parasite prevalence in the host species) may have contributed to the end or prevented the occurrence of A. sanguinea blooms, whereas low parasite prevalence was associated with short-term (≤2 weeks) A. sanguinea blooms. The complete absence of parasitism in 2007 was associated with an extreme A. sanguinea bloom. Anomalously strong upwelling conditions were detected in 2007, suggesting that A. sanguinea was able to outgrow Amoebophrya and ‘escape’ parasitism. We conclude that parasitism can strongly influence dinoflagellate bloom dynamics in upwelling systems. Moreover, Amoebophrya may indirectly influence net phytoplankton species composition, as species that dominated the net phytoplankton and developed algal blooms never appeared to be infected.  相似文献   

8.
The parasitic dinoflagellate Amoebophrya sp. ex Karlodinium veneficum was used to test two hypotheses: (1) infection of cells decreases with increasing host toxicity and (2) parasitism causes the catabolism of host toxin. To test the first hypothesis, host strains differing in toxin content were inoculated with dinospores of Amoebophrya sp. derived from infected cultures of toxic and non-toxic K. veneficum, with resulting infections assessed following 24-h incubations. Contrary to expectations, infection of K. veneficum by Amoebophrya sp. was positively correlated with host toxicity. To examine the second hypothesis, synchronous infection with >80% of cells being parasitized was induced using a toxic strain of K. veneficum, and total toxin concentration (intracellular plus extracellular levels of KmTX1) was followed over the 3-day infection cycle. Toxin content ml−1 increased with growth of K. veneficum in uninfected control cultures, but declined in infected cultures as the parasite completed its life cycle. On a cellular basis, toxin content of infected and uninfected cultures differed little during the experiment, suggesting that the parasite does not actively catabolise host toxin. Rather, infection appears to promote degradation of toxins via death of host cells and subsequent bacterial activity. Results indicate that Amoebophrya sp. ex K. veneficum has greater potential to impact toxic strains relative to non-toxic host strains in natural systems. Thus, Amoebophrya sp. ex. K. veneficum may limit the occurrence of toxic K. veneficum blooms in marine and estuarine environments, while simultaneously functioning as a pathway for dissipation of host toxin.  相似文献   

9.
10.
Li  R.  & Kennelly  P. J. 《Journal of phycology》2003,39(S1):35-35
Parasitic dinoflagellates of the genus Amoebophrya infect and kill bloom-forming dinoflagellates, including the toxic species Karlodinium micrum . Unlike non-toxic hosts, K. micrum is partially resistant to infection, a trait that may be related to toxin production. Here we tested the hypothesis that parasitism of K. micrum is inversely related to toxin concentration in the culture medium. Time-course studies were conducted to determine the influence of extracted toxin and toxin carrier (methanol) on host growth, parasite prevalence, and parasite load. Results indicate that methanol concentrations below 0.1% have no effect on these variables. When methanol concentration was maintained below 0.1%, extracted toxin equivalent to 100 to 10,000 K. micrum per ml had no effect on host abundance. We are currently analyzing sample to assess the fate of Amoebophrya dinospores when exposed to K. micrum toxin. We will also consider the effect of intracellular host toxin on parasite success, by examining the fate of Amoebophrya dinospores when inoculated to K. micrum cultures that express different levels of toxin production. Understanding the effect of toxins on parasite success will contribute to our knowledge of host-parasite biology and provide insight into the role of dinoflagellate toxins as a defense against parasitism.  相似文献   

11.
The endoparasitic dinoflagellate Amoebophrya ceratii Koeppen occurs in coastal waters of Nova Scotia within cells of two dinoflagellate hosts, a Scrippsiella species (probably S. trochoidea (Stein) Loeb. III) and Dino-physis norvegica Claparede & Lachman. We describe the development of the endoparasitic stage (the trophont) of A. ceratii within host cells using light and electron microscopy. After entry into the host, the trophont grows and expands until most of the host cell is occupied by the parasite. Growth is marked by a proliferation of trophont nuclei and flagella and by the formation of numerous lobes, each of which possesses a characteristic dinoflagellate amphiesma. The mature endoparasitic trophont is recognized at the light microscopic level as a beehive-shaped structure that consists of numerous lobes of the developing motile sporont cells and a mastigocoel cavity containing the sporont flagella.  相似文献   

12.
Preliminary attempts to culture Amoebophrya sp., a parasite of Gymnodinium sanguineum from Chesapeake Bay, indicated that success may be influenced by water quality. To explore that possibility, we determined development time, reproductive output, and infectivity of progeny (i.e. dinospores) for Amoebophyra sp. maintained on G. sanguineum grown in four different culture media. The duration of the parasite's intracellular growth phase showed no significant difference among treatments; however, the time required for completion of multiple parasite generations did, with elapsed time to the middle of the third generation being shorter in nutrient-replete media. Parasites of hosts grown in nutrient-replete medium also produced three to four times more dinospores than those infecting hosts under low-nutrient conditions, with mean values of 380 and 130 dinospores/host, respectively. Dinospore production relative to host biovolume also differed, with peak values of 7.4 per 1,000 microm3 host for nutrient-replete medium and 4.8 per 1,000 microm3 host for nutrient-limited medium. Furthermore, dinospores produced by "high-nutrient" parasites had a higher success rate than those formed by "low-nutrient" parasites. Results suggest that Amoebophrya sp. is well adapted to exploit G. sanguineum populations in nutrient-enriched environments.  相似文献   

13.
Miller JJ  Delwiche CF  Coats DW 《Protist》2012,163(5):720-745
Amoebophrya is a syndinian parasite that kills harmful bloom forming algae. Previously uncharacterized ultrastructural aspects of infection and development were elucidated. The biflagellate dinospore has two mitochondria, electron-dense bodies, striated strips, trichocysts, and a nucleus with peripherally condensed chromatin. After finding an Akashiwo sanguinea host and adhering to its surface, the parasite penetrates the host surface, apparently using a microfilament based motility and electron-dense bodies within a microtubular basket in the process of parasitophorous vacuole membrane formation. After entering the host nucleus, possibly by a similar mechanism used to enter the host cell, the parasite cytosol expanded substantially prior to mitosis. From 12-36 hours mitochondria were inconspicuous but present. Chromatin condensation was variable. By 36 hours post-infection, parasites had multiple nuclei, a microtubule-supported cytopharynx, and were beginning to form a fully internal mastigocoel. By 48 hours, the characteristic "beehive" appearance was apparent with flagella projecting into a fully developed mastigocoel. The cytoplasm contained trichocysts, elongated mitochondria, and nuclei with peripherally condensed chromatin. Although Amoebophrya lacks an apical complex, its electron-dense bodies show functional similarities to apicomplexan rhoptries. Its lack of permanently condensed chromosomes, but compact dinospore chromatin, supports the idea that dinoflagellate permanently condensed chromosomes may be a remnant of a parasitic ancestor with a compact dispersal stage.  相似文献   

14.
We developed a rapid and effective procedure for scanning electron microscopy of three delicate dinoflagellates, Karlodinium micrum, Akashiwo sanguinea, and Heterocapsa niei. Good results were obtained when the specimens were fixed with a modified Párducz’s fixative (2% osmium tetroxide:saturated mercuric chloride = 5:1 v/v) for 10 min, washed in 0.05 M sodium cacodylate trihydrate buffer for 2 min, dehydrated in tert-butanol for 10 min and dried with hexamethyldisilazane in air for 3 min in a fume hood because reagents are very toxic. This method could be completed in 25 min. Compared with other preparative techniques, the present protocol has significant advantages for SEM observation by limiting distortion of delicate specimens and reducing the preparation time.  相似文献   

15.
Chesapeake Bay populations of the red-tide dinoflagellate Gymnodinium sanguineum were regularly infected by the parasitic dinoflagellate Amoebophrya ceratii during the summers of 1988–1991. Infections developed inside the nucleus of G. sanguineum and were always lethal to the host. Parasite generation time was ? 40 h at 23° C, with the intracellular, trophont phase lasting 39.5 ± 0.3 h, and the extracellular, vermiform stage persisting for ? 20 min. Near surface accumulations of G. sanguineum sometimes exceeded 1,000 cells/ml; however, host abundance was relatively low when integrated over the surface mixed layer of each station (mean = 12.2 cells/ml ± 2.96 SE; n = 60). Parasitized hosts were encountered in 75% of the samples where host abundance was ≥ 1 per ml, and epidemic outbreaks (20–40% hosts infected) were observed on several occasions. Epidemic infections were generally located several meters below surface accumulations of G. sanguineum and were always restricted to a narrow region near the pycnocline. Consequently, integrated station values for parasite prevalence were low, with an average 2.7% (± 0.31 SE; n = 60). Parasite induced mortality removed up to 8% of G. sanguineum populations per day, but averaged < 2% of host biomass throughout the Bay. Thus, parasitism by A. ceratii does not appear to be a major factor regulating G. sanguineum bloom in the main stem of Chesapeake Bay.  相似文献   

16.
Small subunit rRNA sequences were amplified from Amoebophrya strains infecting Karlodinium micrum, Gymnodinium instriatum and an unidentified Scrippsiella species in Chesapeake Bay. The alignable parts of the sequences differed from each other and from the previously reported rRNA sequence of the Amoebophrya strain infecting Akashiwo sanguinea in Chesapeake Bay by 4 to 10%. This is a greater degree of difference than sometimes found between sequences from separate genera of free-living dinoflagellates. These sequence differences indicate that the Amoebophrya strains parasitizing dinoflagellates in Chesapeake Bay do not all belong to the same species. In spite of their relative dissimilarity, the sequences do group together into a single clade with high bootstrap support in phylogenetic trees constructed from the sequences.  相似文献   

17.
Characterizing ecological relationships between viruses, bacteria and phytoplankton in the ocean is critical to understanding the ecosystem; however, these relationships are infrequently investigated together. To understand the dynamics of microbial communities and environmental factors in harmful algal blooms (HABs), we examined the environmental factors and microbial communities during Akashiwo sanguinea HABs in the Jangmok coastal waters of South Korea by metagenomics. Specific bacterial species showed complex synergistic and antagonistic relationships with the A. sanguinea bloom. The endoparasitic dinoflagellate Amoebophrya sp. 1 controlled the bloom dynamics and correlated with HAB decline. Among nucleocytoplasmic large DNA viruses (NCLDVs), two Pandoraviruses and six Phycodnaviruses were strongly and positively correlated with the HABs. Operational taxonomic units of microbial communities and environmental factors associated with A. sanguinea were visualized by network analysis: A. sanguineaAmoebophrya sp. 1 (r = .59, time lag: 2 days) and A. sanguineaEctocarpus siliculosus virus 1 in Phycodnaviridae (0.50, 4 days) relationships showed close associations. The relationship between A. sanguinea and dissolved inorganic phosphorus relationship also showed a very close correlation (0.74, 0 day). Microbial communities and the environment changed dynamically during the A. sanguinea bloom, and the rapid turnover of microorganisms responded to ecological interactions. A. sanguinea bloom dramatically changes the environments by exuding dissolved carbohydrates via autotrophic processes, followed by changes in microbial communities involving host‐specific viruses, bacteria and parasitoids. Thus, the microbial communities in HAB are composed of various organisms that interact in a complex manner.  相似文献   

18.
Amoebophrya is a parasitic, syndinian dinoflagellate genus that must infect another host dinoflagellate in order to reproduce. Work by Park et al. [Mar. Ecol. Prog. Ser., 227: 281–292 (2002)] has led to the hypothesis that Amoebophrya's development within a host cell nucleus disrupts the flow of genetic information involved in plastidial function. The possibility that genetic disruption by this parasite could lead to alterations in plastidial lipid composition during the course of an infection has not yet been elucidated. Our primary objective in this lipidomic study was to examine the chloroplast membrane galactolipid composition of Alexandrium tamarense infected by an Amoebophrya species in order to determine whether infection of A. tamarense causes a phenotypic alteration in the composition of mono- and digalactosyldiacylglycerol (MGDG and DGDG, respectively), two galactolipids that comprise the majority of photosynthetic membranes. Our secondary objective was to determine if non-photosynthetic Amoebophrya sp. either incorporated host cell MGDG and DGDG, and/or itself produced forms of MGDG and DGDG, as has been observed previously in heterotrophic apicomplexan parasites distantly related to Amoebophrya. We found that, despite development of Amoebophrya sp. within the nucleus, the composition of A. tamarense MGDG and DGDG did not change throughout the infection process. The predominant forms of these galactolipids were 18:5/18:4 (sn???1/sn?2) and 20:5/18:4 DGDG, which were present at similar abundances in both an uninfected host and a host late in the infection process just prior to release of Amoebophrya sp. dinospores. Amoebophrya sp. did not possess appreciable amounts of any forms of MGDG and DGDG.  相似文献   

19.
ABSTRACT. Parasitic dinoflagellates of the genus Amoebophrya play important roles in the ecology of estuaries and open ocean environments. Little is known of the cell and molecular biology of Amoebophrya , but the genus is intermediate on phylogenetic trees between apicomplexans and typical dinophycean dinoflagellates. Here, we constructed four cDNA libraries, from different stages after infecting the host, Karlodinium veneficum , with Amoebophrya sp. These libraries were used to generate 898 expressed sequence tags (ESTs), with sequences attributed to either the host or parasite, based on AT bias, codon usage, and occurrence during infection. Overall, 209 sequences were attributable to the parasite and 685 to the host. The 50 putative parasite sequences with good protein matches in GenBank were used to find the same protein from host ESTs. For 26 genes, both host and parasite sequences were identified, of which 20 encoded ribosomal proteins. PCR for seven predicted parasite and two host genes were used to confirm attributions. The most common host and parasite ESTs were compared to see if multiple gene copies were present. The host plastocyanin gene had multiple sequence variants, but parasite rps 27 a contained only one polymorphism, likely due to an amplification error. Amplification, cloning, and sequencing of five parasite protein-coding genes suggested that the parasite has a single sequence for each gene, but three host genes were found to have multiple variants. The genome of Amoebophrya sp. infecting K. veneficum appears to have an organization more similar to other eukaryotes than to the tandem gene arrangements found in dinoflagellates.  相似文献   

20.
An external transcribed spacer (ETS) walking PCR technique was developed for the isolation of unknown sequences adjacent to the 18S rDNA. This strategy relied on four "walking primers", which were designed to bind unknown sequences upstream from the 18S rDNA, and a specially programmed series of thermocycles. This method was successful in the isolation of the 5' ETS regions from harmful dinoflagellates, including Alexandrium affine, A. catenella, A. minutum, A. tamarense, and Akashiwo sanguinea. Mono-directional sequencing reactions revealed the PCR products to be 392–962 nucleotides in length, and the 5' ETS in these products were longer than 362 bp. These are the first such sequences available for A. sanguinea and the Alexandrium. In comparisons of the ETS sequences, genetic distance was considerably high within the Alexandrium. Furthermore, the sequences were significantly variable among the different strains of identical species: genetic distance was recorded at 0.0420 for A. tamarense strains and as less than 0.7841 within strains of A. sanguinea. The 5'-start nucleotide of 18S rDNA was variable between the two genera: the five species of Alexandrium contained a T base, and A. sanguinea contained an A base. These results demonstrate the effectiveness of the ETS walking PCR method. This method will be valuable in directional ETS walking from known regions to unknown regions, particularly concerning the boundary sequences of rRNA genes.  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