Potential DMSP-degrading Roseobacter clade dominates endosymbiotic microflora of Pyrodinium bahamense var. compressum (Dinophyceae) in vitro

July 8, 2017 | Autor: Deo Onda | Categoria: Harmful algal blooms, DMSP, Roseobacter, Pyrodinium bahamense, Endosymbiotic Bacteria
Share Embed


Descrição do Produto

Arch Microbiol DOI 10.1007/s00203-015-1133-0

SHORT COMMUNICATION

Potential DMSP‑degrading Roseobacter clade dominates endosymbiotic microflora of Pyrodinium bahamense var. compressum (Dinophyceae) in vitro Deo Florence L. Onda1,2 · Rhodora V. Azanza1 · Arturo O. Lluisma1 

Received: 5 March 2015 / Revised: 18 June 2015 / Accepted: 28 June 2015 © Springer-Verlag Berlin Heidelberg 2015

Abstract  Many aspects of the biology and ecology of the toxic dinoflagellate Pyrodinium bahamense var. compressum are still poorly understood. In this brief note, we present identification of its associated intracellular bacteria or endosymbionts via PCR cloning and 16s rRNA gene sequencing and their localization by confocal microscopy, a first for Pyrodinium. The most frequently observed species in the endosymbiotic microflora were from Roseobacter clade (Alphaproteobacteria, 68 %) and Gilvibacter sediminis (Flavobacteriaceae, 20 %). Roseobacter lineage, the most abundant taxa in this study, is known to be involved in dimethylsulfoniopropionate metabolism which is highly produced in dinoflagellates—a possible strong factor shaping the structure of the associated bacterial community.

Communicated by Erko Stackebrandt. Electronic supplementary material  The online version of this article (doi:10.1007/s00203-015-1133-0) contains supplementary material, which is available to authorized users. * Arturo O. Lluisma [email protected] 1

The Marine Science Institute, University of the Philippines, 1101 Diliman, Quezon City, Philippines

2

Present Address: Takuvik/Quebec‑Ocean/IBIS, Pavillon Eugene Marchand, Université Laval, Québec City, QC G1V0A6, Canada



Keywords  Endosymbiotic bacteria · Dimethylsulfoniopropionate (DMSP) catabolism · Symbiosis · Pyrodinium bahemense var. compressum

Introduction Pyrodinium bahamense var. compressum (Bӧhm) Steidinger, Tester and J.F.R. Taylor is one of the most studied toxic dinoflagellates. However, many aspects of its biology and ecology still remain to be understood including its association with its intracellular bacterial community or endosymbionts (Azanza et al. 2006). Identifying these intracellular bacteria is a crucial step toward the elucidation of their roles in the biology and ecology of dinoflagellates. For example, recent evidence showed that the removal of culturable and possibly most bacteria by antibiotic treatment significantly affected the growth or caused death of Pyrodinium cultures (Santos and Azanza 2012). They have also been implicated in the initiation, maintenance and decline of algal blooms (Doucette 1995). Studies showed that dinoflagellates have specifically structured associated bacterial communities beneficial to the host organism. Previous investigations that identified potential symbionts of P. bahamense var. compressum were limited to culture-dependent methods (Azanza et al. 2006; Chin et al. 2013). Hence, the uncultivable members of the endosymbiotic community and factors shaping community structures have remained to be identified and characterized. In this study, we investigated the genetic diversity of the endocytic bacterial community in Pyrodinium. The localization and spatial distribution of these bacteria were also investigated via confocal laser scanning microscopy (CLSM) using a fast and reproducible sample preparation

13



method. We then inferred potential interactions based on the characteristics of most similar cultured bacterial isolates.

Materials and methods Sample preparation for CLSM Pyrodinium bahamense var. compressum PbcMZRVA042595 isolated from Masinloc Bay, Zambales, Philippines, was used in this study. Samples for CLSM imaging were prepared by fixing 1 mL algal sample with 1 % glutaraldehyde (Sigma-Aldrich) and pelleting at 2500×g for 15 min. Cells were resuspended in 100 μL of 1X phosphate buffer saline (PBS) and transferred to a polycarbonate membrane with a hole in the center mounted in a glass slide. Five microliters (5 μL) of each of 5000× SybrGreen (Invitrogen) and 50 % calcofluor white (Sigma-Aldrich) was added. Observation was done in an inverted CLSM (Carl Zeiss LSM 710, Germany), and images were taken by excitation at 385 nm (cell wall), 488 nm (algal nucleus and bacteria) and 580–720 nm (chloroplast autofluorescence). Cell cleaning and bacterial DNA extraction Fifteen milliliters (15 mL) of cell culture samples was fast-fixed (0.1 % glutaraldehyde), harvested by centrifugation at 2500×g for 15 min, resuspended in 1 mL 1 % lysozyme solution and then incubated at room temperature for another 15 min. Subsequent centrifugation was done at 2300×g (13 min) and 2200×g (11 min) with resuspension in 500 µL 1X PBS in between. Extracellular bacterial contamination was assessed by CLSM and PCR (i.e., resuspending medium) before proceeding. Upon confirmation of efficiency of cleaning, cells were pelleted at 2200×g for 10 min and finally resuspended in 1 mL sterile deionized water (Nanopure). For DNA extraction, 100 μL of the resuspended cells was transferred to a 1.5-mL microfuge tube, and four sterile silica beads were added. The sample was then bead-beaten using a vortex mixer (Disruptor Genie) for 3–4 min. Enzymatic purification was done by adding 3 μL of proteinase K (30 mg mL−1, Macherey– Nagel) with incubation at 55 °C for 30 min and stopped at 85 °C for 40 min. Cells were harvested on the 7th, 12th and 21st day of the batch cultures and were then pooled together after extraction. PCR, cloning and 16s rRNA gene library construction The 16s rRNA genes of the endosymbionts were amplified using the universal primers 27F and 1492R following conditions described by Isenbarger et al. (2008) with

13

Arch Microbiol

annealing temperature at 40 °C. Amplification was done in 25 µL reaction mixtures as described by Onda et al. (2014) but with 5 μL DNA template (suspended in water). After PCR, target amplicons (~1.4 kb) were gel purified using Qiaquick Gel Purification kit (Qiagen) and then stored at −20 °C. High-efficiency transformation Electromax DH10BT1 phage resistant cells (Invitrogen) were used for cloning by electroporation using the GenePulser electroporator (BioRad) at 2.0 kV pulse. Cells were plated on LB medium with 50 μg mL−1 kanamycin at 37 °C overnight. A loopful of the colonies was individually resuspended in 40 μL sterile water and lysed at 95 °C for 15 min. The flanking M13 regions of the vector were amplified to check for correct inserts. Amplicons with expected band size (~1.6 kb amplicons) were gel purified and sequenced using the M13F primer (1st Base Gene, Malaysia). Sequence data processing was carried out in MEGA 5.0 (Tamura et al. 2011), and chimeras were screened in DECIPHER (Wright et al. 2011). Most similar sequences were determined by batch BLASTn search against the NCBI GenBank 16s rRNA reference sequence database, then downloaded and aligned using Muscle (Edgar 2004). Phylogenetic trees were generated via maximum likelihood method (using the Kimura-2 parameter model as determined by model test) with bootstrap analysis using 1000 replicates. Since not all sequences were sequenced in the same orientation, separate trees were built using representative sequences from targeted taxa together with the closest related reference sequences. Sequences generated in this study were deposited in NCBI GenBank database under accession numbers KP453874–KP453973.

Results and discussion Taxonomic diversity of associated bacteria In this study, we removed the extracellular/attached bacteria prior to extraction of DNA from the Pyrodinium cells to focus on the endosymbiotic communities by applying washing protocols (Fig. 1a, b). Results show that despite cleaning, we still obtained higher bacterial diversity than previously reported (Azanza et al. 2006; Chin et al. 2013). A total of 100 clones of screened transformants were analyzed and mostly with 98–99 % sequence similarity to the reference sequences (Table 1, Supplementary Fig. 1a, b). However, some sequences did not return hits with significant scores, consistent with previous studies done on environmental/natural samples (Alavi et al. 2001). The most frequent clones are from class α-proteobacteria, clade Roseobacter. Highly abundant putative species observed include Ruegeria sp. and Roseovarius

Arch Microbiol

Fig. 1  CLSM images of Pyrodinium bahamense var. compressum showing a the extracellular/attached bacteria (green dots); b a cleaned algal cell after the washing protocol; c a cross section of the algal cell

showing the localization of the endosymbiotic (end) bacteria along with the NR nuclear material, OM outer membrane, Th theca, Chl chloroplasts

sp. which are also reported in other toxic dinoflagellates and generally in marine environment (Green et al. 2010; Buchan et al. 2005). The closest matches to our sequences were also species from seawater, sand and marine sediments supporting that our results are not artifacts or contamination. Other sequences are similar to putative biofilmassociated bacteria such as Phaeobacter sp. and Leisingera sp., and a phytoplankton phycosphere-associated Marinovum algicola (formerly Roseobacter algicola) (Green et al. 2010). Also, the abundance of single-genus sequences from Flavobacteriaceae of the species G. sediminis (20 %) in the library is surprising since most of its representative cultures are associated with sediments. Recently, it was observed that breaks in the cell wall are created during the regeneration of temporary cysts in sediments which could possibly provide entry of bacteria from the environment (Onda et al. 2014). There are also some representatives from other Bacteroidetes such as Cyclobacterium sp. Inconsistency in the observed diversity with previous reports was also observed (Azanza et al. 2006; Chin et al. 2013) which may be attributed to the biases in the methods used, the differences in conditions used in maintaining the culture and other unknown factors. Biological or even statistical explanations, resulting in disparity in the datasets, are also possible.

As much as 10 % of fixed carbon is directed to DMSP production by marine phytoplankton (Howard et al. 2008). In this study, one interesting common characteristic of the most abundant bacterial groups is their possible involvement in DMSP decomposition (Table 1, Supplementary Fig. 1a, b). Dimethylsulfoniopropionate is a relatively stable molecule, but also a labile bacterial substrate that can supply up to 15 and 100 % of bacterial carbon (C) and sulfur (S) demand, respectively (Simó et al. 2002). Evidence showed that alphaproteobacteria lineage, which makes up to 68 % of the clones in our library (Table 1), is involved in the metabolism and recycling of DMSP (González et al. 1999). DMSP turnover rate in phytoplankton blooms is thus hastened by bacterial uptake (Pinhassi et al. 2005). Most of the DMSP in seawater is degraded through the demethylation/demethiolation pathway which produces intermediates (e.g., methanethiol, methyl mercaptopropionate) that are significant for the assimilation of C and S. When DMSP-S is not assimilated (i.e., low demand for S), the molecule is cleaved by either the algal or bacterial enzyme DMSP lyase which produces the climate active gas dimethylsulfide (DMS) and acrylate. Although different copies of DMSP lyase genes have been found in several phytoplankton, it has not yet been reported in Pyrodinium. The DMS then diffuses into the atmosphere, or oxidized by reactive oxygen species (ROS) or DMS-oxidizing bacteria in seawater producing the end product dimethylsulfoxide [DMSO, (Sunda et al. 2002)] or tetrathionate (Boden et al. 2010), respectively. These two pathways (demethylation/demethiolation and DMSP cleavage), however, are estimated to only account for 5–10 % of the fate of DMSP. Incidentally, as much as 58 % of cells surveyed via metagenomics from ocean surface were found to possibly participate in the

Bacteria–algae association Phytoplankton species produce organic compounds and metabolites that attract the neighboring bacterial community. Dinoflagellates in particular are significant producers of dimethylsulfoniopropionate (DMSP) that serves as osmolyte and osmoprotectant during salinity fluctuations.

13

13 99 96 92 96

NR029273.1, NR112651.1 NR116682.1 NR044547.1 NR043611.1

Marinovum algicola (now Roseobacter algicola) Marivita byunsanensis Oceanibaculum indicum

 426

 368, 373

97

95

NR113209.1

99 97

NR108232.1 NR109594.1

Roseovarius indicus

95

98 96 92

NR108509.1 NR036802.1 NR074599.1

Ruegeria pomeroyi Ruegeria sp. Tropicibacter multivorans Methylophaga thalassica Pseudomonas protegens Thioalkalivibrio sulfidophilus

 66, 73, 84, 125, 244, 246, 353, 370

 114, 132

 211, 232, 240

Gammaproteobacteria  88

 113, 134

 103, 356, 449

Bacteroidetes  Flavobacteriacea   62, 69, 71, 77, 108, 110, 116, 124, 126, 133, 141, 213, 380, 381, 404, 422, 427, 432, 434, 435  Flexibacteriaceae   91, 119

93

NR074151.1

Ruegeria arenilitoris

99

91

NR114011.1

NR040920.1

Marinoscillum furvescens

88

98

98

Gilvibacter sediminis

NR074692.1

NR074150.1

NR109635.1

99

 70, 92, 204, 236, 362, 371, 376, 377, 379, 445

NR043564.1

Roseovarius litoreus Roseovarius pacificus

 65, 120, 131, 358

 127, 378, 430, 431

94

 231, 237

NR043932.1 NR121734.1

Pseudoruegeria aquimaris Roseibacterium elongatum

95

 450

NR113210.1

 118

Poseidonocella pacifica Poseidonocella sedimentorum

 50

 428

97

NR074144.1

Phaeobacter gallaciensis

98

 364

NR118542.1

Pelagibaca bermudensis Phaeobacter caeuruleus

 369, 420

 17, 49, 80, 89, 111, 352, 410, 410b

90

 36, 52, 83, 86, 93, 112, 117, 212, 242, 363, 412

NR042670.1 NR074160.1

Leisingera aquimarina

96

Maricaulis maris

NR102914.1

Marine organism

Marine sediment

Highly saline bioreactor

Plant symbiont

Marine microbial mat

Seawater

Type strain

Atlantic coastal water

Seashore sand

Deep-sea sediment

Seawater

Deep seawater

Sand

Seawater, Korea

Shallow sandy sediment

Shallow sandy sediment

Coastal environment

Marine environment, Palau and Japan

Sargasso sea

Deep seawater, Indian Ocean

Tidal flat sediment

Marine environment, Spain; phycosphere of Prorocentrum lima

Type strain

Marine electroactive biofilm

Industrial strain

% similarityb (%) Source

 101, 115, 249, 419

Ketogulonicigenium vulgare

Alphaproteobacteria  415

Reference sequence NCBI accession number

 441

Taxonomic groupa

Clones

Table 1  Summary of taxonomic assignments based on sequence similarity search using BLASTn against the NCBI GenBank

Arch Microbiol

Type strain

Marine oilfield sediment, SCS

85

97

  Maximum percentage of identical nucleotides in the noted alignment length

  107

  Full name or description of the closest matched sequence

b

a

Planctomyces maris

NR025327.1

NR043903.1 Cyclobacterium lianum  Cyclobacteriaceae   122, 135, 360

 Planctomycetes

Taxonomic groupa Clones

Table 1  continued

Reference sequence NCBI accession number

% similarityb (%) Source

Arch Microbiol

marine sulfur cycle via the DMSP catabolic pathway (Howard et al. 2008). Several isolates of Roseobacter in particular were observed to demethylate DMSP in culture. The corresponding genes for DMSP catabolism were also found in some of the reference type strains in this study such as Ruegeria pomeroyi DSS-3 and Phaeobacter gallaeciensis DSM 17395. Although species belonging to Roseobacter clade are cosmopolitan in nature, they have been frequently and abundantly found to be associated with DMSP-producing dinoflagellates such as Pfiesteria sp. (Alavi et al. 2001), Gymnodinium catenatum (Lafay and Ruimy 1995, Green et al. 2004), Alexandrium sp., Scripsiella sp. (Hold et al. 2001), Prorocentrum lima (Lafay and Ruimy 1995), A. tamarense (Jasti et al. 2005) and Pyrodinium bahamense var. compressum (this study) among others. These studies suggest how conditions inside the cell could possibly shape the structure of the associated bacterial communities which could have beneficial role to the host organism. For example, a symbiont Ruegeria sp. TM1040 which is capable of DMSP demethylation was isolated from a culture of the heterotrophic dinoflagellate Pfisteria piscicida (Miller and Bellas 2004). Further studies later suggested that the association is obligate and is affected by changing conditions. Other derivatives from DMSP decomposition may also be reabsorbed by the alga or used up by other bacteria. Methylophaga thalassica, for example (Supplementary Fig. 1b), an obligate methylotrophic bacteria found in environments with high DMS (Boden et al. 2010), was also observed to be present in Pyrodinium. The dominance of potential DMSP catabolic clades inside the Pyrodinium cell raises a lot of other questions that may help further understand DMSP regulation and DMS production in dinoflagellates. Further investigations using more laboratory-based approaches, however, are still necessary to provide evidence for such hypothesis. Confocal laser scanning microscopy images also showed that the associated bacteria seem to be more abundant and interspersed together with the chloroplasts in the cytoplasm in very close proximity with the theca (Fig. 1c). It has been shown that Roseobacter clade uniquely produces bacteriochlorophyll a with some species capable of aerobic anoxygenic photosynthesis (Allgaier et al. 2003). Recent reports demonstrated that DMSP uptake of Dinoroseobacter shibae associated with Prorocentrum minimum could be regulated by the absence or presence of light. Depending on necessity, light availability and aerobic anoxygenic photosynthetic activity could also influence shift from mutualistic to parasitic relationships between the host and the symbionts (Wang et al. 2014). Further studies are needed, however, to show that the physical closeness of chloroplasts to the bacteria in Pyrodinium implies possible functional relationships. Other taxa such as Bacteroidetes could also supply iron and vitamins, recycle organic compounds and

13



metabolize toxins (Kodama et al. 2006). Gilvibacter sediminis which consists 20 % of the clones in our library also singly dominated the Flavobacteriaceae group. It has been previously reported in the benthic dinoflagellate Coolia monotis (Ruh et al. 2009), in sediments (Khan et al. 2007) and as free-living bacterioplankton (Brettar et al. 2012). Interestingly, although most Flavobacteria have been generally implicated in the conversion of high molecular weight substrates to smaller components (Buchan et al. 2014), the particular role potentially played by G. sediminis in associations has not yet been fully explored. The contribution of the associated bacterial species to algal toxicity, specifically saxitoxin (STX), has also been suggested (Azanza et al. 2006; Hold et al. 2001). Gene homologs most similar to bacteria that could possibly be involved in the STX biosynthetic pathway in Pyrodinium have also been observed (Onda et al., unpublished data). Although phylogenetic data do not directly suggest functionality in the bacterial community, it could still provide insights into the functional roles of the endosymbionts and how groups correlate to certain variables significant to the host organism. The high diversity of intracellular bacteria found in this study, for example, suggests strong environmental conditions that favored such associations to exist which in turn could have ecological implications. These associations and their still unknown interactions may be truly significant that their removal from the host and/or disruption of the association may be detrimental to the host alga. For example, loss of viability of cultured Pyrodinium was observed when associated cultivable bacteria were attempted to be removed by antibiotic treatment (Santos and Azanza 2012). These results are the foundations of our ongoing investigations to further understand the significance of the “microalgal microbiome” in such ecological relationships using high-throughput sequencing approach. Pyrodinium’s relationship with its associated bacterial community is complex and still enigmatic. This model system, however, provides a wealth of information for further understanding coevolution and host–symbiont interactions. These symbionts are also known producers of secondary metabolites—a mine for potential novel genes with biotechnological applications. Acknowledgments  This study was part of the research program “Ecology and Oceanography of Harmful Algal Blooms in the Philippines (PhilHABs), Project 1: Biodiversity/Genetic Diversity of selected HAB-forming species in the Philippines and their associated bacterial communities”, funded and supported by the Department of Science and Technology (DOST) through the Philippine Council for Aquatic and Marine Research and Development (PCAMRD, now part of PCAARRD). We would also like to thank the Biotechnology and Molecular Genetics Research Laboratory at the UP-NIMBB headed by Dr. Ameurfina D. Santos for the use of the electroporator, Dr. Neda Barghi for her assistance in cloning, Emelita Eugenio for her help in

13

Arch Microbiol maintaining the algal cultures, Dr. Mary Anne Santos for her comments on Pyrodinium biology, and Margaux Goudal for her valuable insights on DMSP regulation and metabolism. Compliance with Ethical Standards  Conflict of interest  The authors express no conflict of interest in publishing this article.

References Alavi M, Miller T, Erlandson K, Schneider R, Belas R (2001) Bacterial community associated with Pfiesteria-like dinoflagellate culture. Environ Microbiol 3:380–396 Allgaier M, Uphoff H, Felske A, Wagner-Dobler I (2003) Aerobic anoxygenic photosynthesis in Roseobacter clade bacteria from diverse marine habitats. Appl Environ Microbiol 69:5051–5059 Azanza MP, Azanza RV, Vargas VMD, Hedreyda CT (2006) Bacterial endosymbionts of Pyrodinium bahamense var. compressum. Microb Ecol 52:756–764 Boden R, Kelly DP, Murrell JC, Schäfer H (2010) Oxidation of dimethylsulfide to tetrathionate by Methylophaga thiooxidans sp. nov.: a new link in the sulfur cycle. Environ Microbiol 12:2688–2699 Brettar I, Christen R, Hofle MG (2012) Analysis of bacterial core communities in the central Baltic by comparative RNA-DNAbased fingerprinting provides links to structure-function relationships. ISME J 6:195–212 Buchan A, González JM, Moran MA (2005) An overview of the marine Roseobacter lineage. Appl Environ Microbiol 70:2560–2565 Buchan A, LeCleir GR, Gulvik CA, González JM (2014) Master recyclers: features and functions of bacteria associated with phytoplankton blooms. Nat Rev Microbiol 12:686–698 Chin GJWL, Teoh PL, Kumar SV, Anton A (2013) Ribosomal DNA analysis of marine microbes associated with toxin-producing Pyrodinium bahamense var. compressum (Bohm), a harmful algal bloom species. Pertan J Trop Agric Sci 36:179–188 Doucette GJ (1995) Assessment of the interaction of prokaryotic cells with harmful algal species. In: Lassus P, Arzul G, Erard-Le Denn E, Gentien P, Marcaillou-Le Baut C (eds) Toxic marine phytoplankton. Lavoisier, Paris, pp 385–394 Edgar RC (2004) MUSCLE: a multiple sequence alignment method with reduced time and space complexity. BMC Bioinformatics 5:113 González JM, Kiene RP, Moran MA (1999) Transformation of sulfur compounds by an abundant lineage of marine bacteria in the alpha-subclass of the class Proteobacteria. Appl Environ Microbiol 65:3810–3819 Green DH, Llewellyn LE, Negri AP, Blackburn SI, Bolch CJS (2004) Phylogenetic and functional diversity of the cultivable bacteria associated with the paralytic shellfish poisoning dinoflagellate Gymnodinium catenatum. FEMS Microb Ecol 47:345–357 Green DH, Hart MC, Blackburn SI, Bolch CJS (2010) Bacterial diversity of Gymnodinium catenatum and its relationship to dinoflagellate toxicity. Aquat Microb Ecol 61:73–87 Hold GL, Smith EA, Rappe MS, Maas EW, Moore RB, Stroempl C, Stephen JR, Prosser JI, Birkbeck H, Gallacher S (2001) Characterization of bacterial communities associated with toxic and non-toxic dinoflagellates: Alexandrium spp. and Scrippsiella trochoidea. FEMS Microbiol Ecol 37:161–173 Howard EC, Sun S, Biers EJ, Moran MA (2008) Abundant and diverse bacteria involved in DMSP degradation in marine surface waters. Environ Microbiol 10:2397–2410

Arch Microbiol Isenbarger TA, Finney M, Rios-Velasquez C, Handelsman J, Ruvkun G (2008) Miniprimer PCR: a new lens for viewing the microbial world. Appl Environ Microbiol 74:840–849 Jasti S, Sieracki ME, Poulton NJ, Giewat MW, Rooney-Varga JN (2005) Phylogenetic diversity and specificity of bacteria closely associated with Alexandrium spp. and other phytoplankton. Appl Environ Microbiol 71:3483–3494 Khan ST, Nakagawa Y, Harayama S (2007) Sediminibacter furfurosos gen. nov., sp. nov. and Gilvibacter sediminis gen. nov., sp. nov., novel members of the family Flavobacteriaceae. Int J Syst Evol Microbiol 57:265–269 Kodama M, Doucette GK, Green DH (2006) Relationships between bacteria and harmful algae. Spinger, Berlin, pp 243–255 Lafay B, Ruimy R (1995) Rausch De Traubenberg C. Roseobacter algicola sp. nov., a new marine bacterium isolated from the phycosphere of the toxin-producing dinoflagellate Prorocentrum lima. Int J Syst Bacteriol 45:290–296 Miller TR, Bellas R (2004) Dimethylsulfoniopropionate metabolism by Pfisteria-associated Roseobacter spp. Appl Environ Microbiol 70:3383–3391 Onda DFL, Lluisma AO, Azanza RV (2014) Development, morphological characteristics and viability of temporary cysts of Pyrodiniumbahamense var. compressum (Dinophyceae) in vitro. Eur J Phycol 49:265–275 Pinhassi J, Simó R, González JM, Vila M, Alonso-Sáez L, Kiene RP, Moran MA, Pedrós-Alió C (2005) Dimethylsulfoniopropionate turnover is linked to the composition and dynamics of the bacterioplankton assemblage during a microcosm phytoplankton bloom. Appl Environ Microbiol 71:7650–7660

Ruh WW, Ahmad A, Mat Isa MN, Mahadi NM, Marasan NA, Usup G (2009) Diversity of bacteria associated with the benthic marine dinoflagellates Coolia monotis and Ostreopsis ovate from Malaysian waters. J Sci Tech Trop 5:23–33 Santos MAG, Azanza RV (2012) Responses of Pyrodinium bahamense var. compressum and associated cultivable bacteria to antibiotic treatment. J Appl Phycol. doi:10.1007/s10811-011-9701-4 Simó R, Archer SD, Pedrós-Alió C, Gilpin L, Stelfox-Widdicombe CE (2002) Coupled dynamics of dimethylsulfoniopropionate and dimethylsulfide cycling and the microbial food web in surface waters of the North Atlantic. Limnol Oceanogr 47:53–61 Sunda WD, Kieber WDJ, Kiene RP, Hunstman S (2002) An antioxidant function for DMSP and DMS in marine algae. Nature 418:317–320 Tamura K, Peterson D, Peterson N, Stecher G, Nei M, Kumar S (2011) MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol Biol Evol 28:2731–2739 Wang H, Tomasch J, Jarek M, Wagner-Dӧbler IW (2014) A dualspecies co-cultivation system to study the interactions between Roseobacters and dinoflagellates. Front Microbiol. doi:10.3389/ fmicb.2014.00311 Wright ES, Yilmaz S, Noguera DR (2011) DECIPHER, search-based approach to chimera identification for 16s rRNA sequences. Appl Environ Microbiol 78:717–725

13

Lihat lebih banyak...

Comentários

Copyright © 2017 DADOSPDF Inc.