Skip to main content
  • Extended genome report
  • Open access
  • Published:

Complete genome sequence and genomic characterization of Microcystis panniformis FACHB 1757 by third-generation sequencing

Abstract

The cyanobacterial genus Microcystis is well known as the main group that forms harmful blooms in water. A strain of Microcystis, M. panniformis FACHB1757, was isolated from Meiliang Bay of Lake Taihu in August 2011. The whole genome was sequenced using PacBio RS II sequencer with 48-fold coverage. The complete genome sequence with no gaps contained a 5,686,839 bp chromosome and a 38,683 bp plasmid, which coded for 6,519 and 49 proteins, respectively. Comparison with strains of M. aeruginosa and some other water bloom-forming cyanobacterial species revealed large-scale structure rearrangement and length variation at the genome level along with 36 genomic islands annotated genome-wide, which demonstrates high plasticity of the M. panniformis FACHB1757 genome and reveals that Microcystis has a flexible genome evolution.

Introduction

The massive development of bloom-forming cyanobacteria is causing problems in eutrophic water bodies worldwide. Among the cyanobacteria, Microcystis is perhaps the most notorious. Many Microcystis species have been reported to be able to produce microcystins [14], which threaten many aquatic ecosystems and cause serious and occasionally fatal human liver, digestive, neurological, and skin diseases [57].

Microcystis is a genus of unicellular colony-forming cyanobacteria whose taxonomy is still unclear [8]. Although morphological criteria have been proposed to distinguish Microcystis species from field samples, such criteria have long been questioned for use in species identification within the genus [9]. Several studies attempted to reconcile molecular and morphological taxonomy in Microcystis [914], and a morphology-based taxonomic system has been dominantly used. Microcystis panniformis was first reported in 2002 and was morphologically described as having flattened, irregular, monolayer colonies with small holes inside and later disintegrated into small pieces [15]. Since the M. panniformis strain SPC 702 was successfully isolated from Lago das Garças, São Paulo in 1999, studies addressing different aspects of this species have been performed [1625]. In China, M. panniformis was reported as a newly recorded species in 2012 [26], and one strain (FACHB1757) was isolated from Lake Taihu. Microcystis panniformis was originally thought to only be distributed in tropical regions, but we showed that this species has invaded the subtropical regions with a monsoon climate [26]. Global expansion of harmful cyanobacteria has been thought to be linked to climate changes, particularly increasing amounts of atmospheric CO2 and surface temperature, which may promote Microcystis growth and enhance the potential for bloom occurrence [2729]. Therefore, a deeper understanding of the ecology and physiology of M. panniformis FACHB1757 by obtaining a robust genome reference may provide insight into the expansion and invasion mechanisms of Microcystis .

Organism information

Classification and features

A water bloom sample was collected directly from the water surface using a plastic bucket in Meiliang Bay of Lake Taihu in August 2011 (Fig. 1a). Lake Taihu (E 30°56′–31°33′,N 119°54′–120°36′), the third largest freshwater lake in China, is located in the south of the Yangtze River Delta. The total area of the lake is 2338 km2, with an average depth of 2 m and total capacity of 47.6 × 108 m3. Lake Taihu is situated in the subtropical zone with a humid and semi-humid monsoon climate, and has suffered from severe eutrophication over the past three decades. Meiliang Bay is located in the northern part of Lake Taihu (Fig. 1a), which has a surface area of 100 km2, depth of 1.8–2.3 m, and is currently hypereutrophic [30].

Fig. 1
figure 1

Strain collection location and photomicrographs of M. panniformis FACHB1757. The strain was originally isolated from Meiliang Bay of Lake Taihu in August 2011 and deposited in the Freshwater Algae Culture Collection at the Institute of Hydrobiology (FACHB-collection, China) with the unique identifier FACHB1757 in 2012. a The precise position of the isolated sample is indicated by a star; WT means Wutang station in Lake Taihu. b The morphology of the strain colonies in the white disk, which were collected directly from the water surface using a plastic bucket (on September 15, 2013 in Meiliang Bay, photo with a Nikon D7000). c, d Flat colonies with small holes as viewed under an optical microscope

Some Microcystis colonies in the sample disintegrated during the sample collection process; thus, only those macroscopic colonies with significant monolayer were collected with 3-ml pipets (BD Falcon, USA), and transferred into 50-ml centrifuge tubes (Corning, USA), and immediately shipped to the laboratory. Finally, macroscopic colonies that had flattened irregular up to monolayers with small holes (in old colonies) were identified as M. panniformis by examination under an optical microscope. Microcystis panniformis FACHB1757 was obtained, and this strain was then stored at the Freshwater Algae Culture Collection at the Institute of Hydrobiology, Chinese Academy of Sciences.

The general characteristics of M. panniformis FACHB1757 are summarized in Table 1, and a phylogenetic tree based on 16S rRNA sequences is shown in Fig. 2. The spherical cells are estimated with a diameter of 2.6 to 6.8 μm (mean 4.7 μm), be densely agglomerated, and form irregular colonies with small holes. The young stages formed small clusters of cells, which were flat or circular in outline, sometimes spheroidal, and with or without an internal hollow. The old stages formed colonies with small holes, which later disintegrated into small groups. The mucilage (margin of colonies) was diffuse, and cells did not overlap. The margin of the colonies was smooth or (in old colonies) irregular. Cell density was regular and evenly agglomerated, sometimes in indistinct rows. Diagnostic characteristics included flat colonies with small holes, toxicity, homogeneously arranged cells, and life cycle was characterized by distinct benthic and planktonic phases [15, 31]. The distribution was tropical, and this is likely a pantropical species (e.g., S. Africa, N. Australia, S. America, Africa, China, Vietnam and New Zealand) [13, 15, 16, 26, 31, 32].

Table 1 Classification and general features of M. panniformis FACHB1757 according to the MIGS recommendations [69]
Fig. 2
figure 2

Phylogenetic tree showing the position of M. panniformis FACHB1757. The dendrogram is based on the 16S ribosome RNA complete sequence of M. panniformis FACHB1757, M. aeruginosa NIES843, M. aeruginosa PCC7806, M. aeruginosa NIES 2549, and representatives of other cyanobacterial genera (Synechocystis, Pseudanabaena, Synechococcus, Thermosynechococcus, Planktothrix, Dolichospermum, Anabaena, Cylindrospermopsis, Nodularia, Nostoc, Aphanizomenon, Raphidiopsis) downloaded from NCBI (sequences without accession numbers were extracted from annotation files of the corresponding genomes) using the neighbor-joining algorithm with 100 bootstrap replications using MEGA6. A bootstrap consensus tree was constructed and is shown. The two copies of 16S rRNA sequences of M. panniformis FACHB1757 are labeled in red. The relationship between M. panniformis FACHB1757 and other important algae species in Cyanophyceae are demonstrated. Species colored in green have whole genome data available in NCBI

Phylogenetic analysis

Whole genome comparative analysis between M. panniformis FACHB1757 and 13 other cyanobacterial species was performed. General information of related genome data is shown in Table S1 (Additional file 1), and all data sets were downloaded from NCBI. The main water bloom-forming cyanobacterial species in freshwater and brackish water worldwide, particularly those in the Lake Taihu region, were included. Unicellular colony-forming Microcystis and filamentous heterocystous Dolichospermum (formerly known as the planktonic Anabaena ) were the main components of cyanobacterial blooms in Lake Taihu [33]. The Aphanizomenon , Pseudanabaena , Cylindrospermopsis , Raphidiopsis, Planktothrix , Synechocystis , and Synechococcus species occurred as dominant species or accompanying species in blooms of Lake Taihu (including Lake Wuli) across different seasons. Among the 14 genome sequences, 691 single-copy gene families were annotated by OrthoMCL (version 2.0.9) [34], and MEGA6 [35] was used to construct a phylogenetic tree based on these sequences (Fig. 3).

Fig. 3
figure 3

Phylogenetic tree of water bloom-forming cyanobacterial species and representative cyanobacteria. The nucleotide divergence tree was constructed using the neighbor-joining algorithm based on 691 sequences of single-copy gene families annotated by OrthoMCL with 100 bootstrap replicates. The representative cyanobacteria that cannot form water blooms are indicated with an asterisk

The phylogenetic tree shows that M. panniformis FACHB1757 and M. aeruginosa NIES843 shared a significantly high similarity, and there was no clear division between M. panniformis and M. aeruginosa strains in the phylogenetic tree. The Microcystis lineage is distinct from the lineage that contains the unicellular Synechocystis , Synechococcus , and other multicellular cyanobacteria. Furthermore, the Synechocystis sp. PCC 6803 genome is more closely related to Microcystis than other strains. This result is congruent with previously published results based on 16S rRNA sequences [3639]. Topological relationships between species in the phylogenetic tree based on single-copy gene families were generally consistent with the phylogenetic tree based on 16S rRNA sequences (Fig. 2).

Although Microcystis can be identified based on 16S rRNA and single-copy gene families sequences at the genus level, taxonomy of Microcystis at the species level was controversial in the past few decades, and five species have even been unified into a single species [13]. 16S rRNA sequence estimation can be ambiguous when analyzing certain Microcystis species with distinct morphologies, as occurred when analyzing M. panniformis and M. ichthyobabe (Fig. 2). Therefore, the whole reference genome sequence data was expected to play a crucial role in species classification of Microcystis . However, the currently available cyanobacterial genome sequences are highly limited. Only three Microcystis strains with complete genomic sequences are available, including M. aeruginosa NIES843 and NIES2549, and M. panniformis FACHB1757 reported here. Furthermore, the further species concepts and more useful molecular approaches should be proposed to classify the species/strain divergences in Microcystis [40, 41].

Genome sequencing information

Genome project history

Microcystis panniformis FACHB1757 was selected for sequencing because of its obvious morphological characteristics; in particular, the macroscopic colonies with significant monolayer can even exceed 30 mm during the summer and early autumn in Lake Taihu. More importantly, until recently, only complete genomes of M. aeruginosa strains (including strains NIES843 and NIES2549) have been published. The complete genome sequence of M. panniformis FACHB1757 would only be the third for Microcystis . The sample information for M. panniformis FACHB1757 is available in NCBI under BioSample ID SAMN03392520. A DNA library with an insert size of 10 Kb was constructed, and the whole genome was sequenced to 48-fold coverage. The completed genome sequence was assembled and uploaded to GenBank under accession number. CP011339. Project details were deposited to NCBI BioProject PRJNA277430. A summary of the project information can be found in Table 2.

Table 2 Project information

Growth conditions and genomic DNA preparation

Microcystis panniformis FACHB1757 colonies collected from the field were grown in MA medium [42] and incubated in 24-well culture plates for 4 wk. Then, floating colonies were transferred to the capped tubes that contained 5 ml of MA culture medium to finally form a unialgal culture. All cultures were grown at 25 ± 1 °C with a 12 h light/12 h dark cycle under a photon irradiance of 25 μmol photons/(m2 · sec) provided by daylight fluorescent lamps. Total genomic DNA of M. panniformis FACHB1757 was extracted using a commercial DNA isolation kit (DNeasy® Plant Mini Kit, Qiagen, USA) following the manufacturer’s instructions, and analyzed by micro-volume fluorescence detection (NanoDrop 8000 Spectrophotometer, Thermo Scientific, USA) and electrophoresis in 0.8 % agarose gel stained with ethidium bromide. The isolated DNA was eluted with 50 μl of the elution buffer from the commercial kit and then stored at −20 °C until subsequent analyses.

Genome sequencing and assembly

First, the genome was surveyed using an Illumina Hiseq sequencer to detect the purity of the cultured unialgal strain. The insert size of the next generation pair-end library was 100 bp, and 1 Gbp raw data was produced in total. All reads were mapped to the M. aeruginosa NIES843 reference complete genome, and more than 80 % of reads matched well. Subsequently, the genome was sequenced using PacBio RS II. Genomic DNA was sheared by Covaris S220 g-TUBE. A 10 Kb library was constructed using a PacBio template prep kit and sequenced using the PacBio SMRT platform. In total, two SMRT cells were run, and 303 megabase pair raw data was obtained. After filtering, the mean read length was 7143 bp with a quality of 0.84, and the longest read was 31,225 bp. HGAP (version 2.2.3) was used for genome assembly. Long reads were chosen as seeds, and the other reads were mapped to the seeds using Blasr (version 1.3.1.132871) [43] for error correction. After alignment, the accuracy of seed sequences were optimized to 99 % to meet the requirements of the Sanger assembly software. There was a total of 128 Mbp of high quality long seed reads, which had an average length of 7898 bp. Celera Assembler (version 8.1) [44] was then used to assemble the seed reads into contigs and Quiver [45] was used for second error correction. Contigs were assembled into the final complete genome sequence using minimus2 in AMOS (version 3.1.0). The final genome consisted of a complete circular 5,686,839 bp chromosome with a GC content of 42.35 % and a 38,683 bp plasmid with a 43.97 % GC content. Sequencing depths were 44.85 and 128.42, respectively.

Genome annotation

TRs were predicted by Tandem Repeat Finder (version 4.07b) [46] and Microsatellite identification tool (version 1.0), which can both identify perfect and compound micro-satellites. Prediction and annotation of the genome were done using the RAST server (version 2.0) [47]. RAST integrated tRNAscan-SE, and the search_for_rnas tool was used to call RNA genes across the chromosome. For gene estimation, GLIMMER2 was used to represent putative genes. Subsequently, a similar search was performed against FIGfams to identify the determined genes and annotate their functions. Moreover, all putative protein-coding genes were assigned to a category using databases including Clusters of Orthologous Groups (COG), Gene Ontology (GO), Kyoto Encyclopedia of Genes and Genomes (KEGG), Swiss-Prot, and Non-Redundant Protein Database.

Genome properties

The genome assembly contained a complete circular chromosome sequence (5.69 M) and a plasmid (38.68 K). The schematic representation of the circular chromosome of M. panniformis FACHB1757 was showed in Fig. 4. Related genome assembly and annotation information can be found in Table 3. Nucleotide homology search of M. panniformis FACHB1757 and M. aeruginosa NIES843 genomes was conducted by BLAST, and similarity between the two species was 83.82 % (Additional file 1: Figure S1). A total of 1944 TRs were found in the genome, including 27 microsatellites, 1742 mini-satellites, and 176 satellites. Genome statistics are shown in Table 4. In total, there were 6567 genes, which included 48 RNA genes and 6519 protein-coding genes. Among the 6519 proteins, most contained around 100 amino acids (Additional file 1: Figure S2), and by compared with function databases mentioned above, 60.15 % of them were determined to have specific functions. There were 42 tRNA genes, and two copies of the rRNA gene cluster were found in the same direction. Function assignments of 6519 putative protein-coding genes were searched against several frequently used databases mentioned above; 3260 genes were assigned to COGs, of which 235 participated in signal transduction. Search of Pfam omains detected 3997 candidates. According to the subsystem classification results processed by RAST, 72 % of determined genes belong to specific subsystems, and the distribution of each category is demonstrated in (Additional file 1: Figure S3). The result of COG function annotation is shown in Table 4, and details of each COG cluster can be found in Additional file 2. The genes assigned to GO categories by InterProScane (version 5.4-47.0) [48] were classified into cellular components, molecular functions, and biological processes. Genes distributed in each category and their functions are shown in (Additional file 1: Figure S4). In the GO data, 309 signal function-related genes were found. KEGG matched 897 functional genes to related systems, as shown in (Additional file 1: Figure S5). Final gross function annotation outcomes are provided in (Additional file 1: Table S2).

Fig. 4
figure 4

Schematic representation of the circular chromosome of M. panniformis FACHB1757. The scales indicate location in Mbp, starting with the initial coding region. Using Circos integrated the gene prediction results of COG function annotation, methylated modification, and some other information. From inner to outer circles: the first circle shows the GC skew (in purple and green), and the value is plotted as the deviation from the average GC skew of the entire chromosome sequence. The bars in the second circle (in black and red) represent the GC content, which is plotted using a 10-Kb sliding window. Positions of tRNA and rRNA are marked by green bars in the third circle. Bars in the fourth and fifth circle are colored according to COG function categories of CDS; the fourth is a backward strand and fifth is a forward strand. The sixth and seventh circles indicate m4C and m6A sites in CDS/rRNA/tRNA regions (in blue bars); the sixth circle is a backward strand and the seventh circle is a forward strand. In the eighth circle, red bars show the m4C and m6A sites in intergenic regions

Table 3 Genome statistics
Table 4 Number of genes associated with general COG functional categories

Insights from the genome sequence

Comparative Microcystis species genomes

Gene ortholog analysis

Genes of four Microcystis species were compared (Fig. 5), and 2669 highly conserved orthologous genes were shared, which are representative of the core genome. Moreover, each genome had strain-specific genes, which varied from 296 to 1900. The M. aeruginosa NIES2549 genome, which has 1388 unique genes, is 1.5 Mbp smaller than that of M. aeruginosa NIES843, which only has 296 unique genes ( M. aeruginosa NIES843 has 1388). Microcystis panniformis FACHB1757 was shown to have 1900 specific genes, which was the greatest amount among the four strains, even though its genome was not the longest.

Fig. 5
figure 5

Venn diagram of gene numbers of four Microcystis species. Less than half of all genes were found in all four species

Secondary metabolite gene clusters

Microcystin was reported to enhance colony formation in Microcystis spp. and plays a key role in the persistence of their colonies and the dominance of Microcystis [49]. As in M. aeruginosa NIES843 and M. aeruginosa PCC7806 genomes, the microcystin synthetase gene cluster (mcyA-J) was highly conserved in M. panniformis FACHB1757 from coordinates 3,496,704 to 3,541,027. Additionally, the distinct thioesterase type II coding gene mcyT, which occurs in toxic strains, and 4-PPT transferase (4-PPTase) were both located far from the mcy gene cluster at coordinates 869,702 to 869,286 and 915,377 to 916,039, respectively, which are similar to the distributions observed in M. aeruginosa NIES843. Notably, there was an absence of mcnA and mcnB in the M. panniformis FACHB1757 chromosome. mcnA codes polyketide biosynthesis proteins, and mcnB is the first open reading frame of mamestra configurata nucleopolyhedrovirus B. Together with mcnC and mcnE, these four genes compose the cyanopeptolin synthesis gene cluster. mcnD was not found in the M. panniformis FACHB1757 genome; thus, the cyanopeptolin produced was non-halogenated and identical to that of M. aeruginosa NIES843 and PCC7806. Toxins may contribute to the adaptation of this strain to its specific ecological niche in eutrophic waters of tropical and subtropical zones. In addition, a putative polyketide synthase gene cluster, which may encode additional small polypeptides found in M. aeruginosa NIES843 (coordinates 2,508,556–2,513,289), was detected in M. panniformis FACHB1757 at coordinates 4,425,371 to 4,430,104. The change in location of the genes mentioned above reflected the extensive structural variation between M. panniformis FACHB1757 and M. aeruginosa NIES843.

Conserved gene clusters

Four functional clusters of conserved genes related to microcystin synthesis, colony formation, photo-regulation, and nutrient assimilation were also compared among the four Microcystis strains. In the microcystin synthesis gene cluster, the mcy and mcn gene clusters were not found in M. aeruginosa NIES2549. This is consistent with the results of a previous study, which showed that M. aeruginosa NIES2549 is a nontoxic strain [50]. With regard to colony formation, M. aeruginosa , M. wesenbergii, and M. panniformis all have typical macroscopic colony structure when observed by naked eye in Lake Taihu during summer and autumn water blooms. Microcystis panniformis seems to be the largest, and can even have more than 30 mm colonies. Polysaccharides and microcystin play important roles in the process of Microcystis colony formation. The maximum EPS content was found in M. wesenbergii and M. aeruginosa , which are not the largest and are only approximately 100 μm [51], but positive correlations between EPS and Microcystis colony size in cultures were supported by previous studies [5254]. mrpC and epsL were absent from all four strains, and only M. aeruginosa NIES843 contained cpsF, although tagH, capD, csaB, and rfbB were conserved in all four strains. Furthermore, mvn codes for a lectin in M. panniformis FACHB1757 and M. aeruginosa PCC7806, which specifically binds to a sugar moiety present on the surface of Microcystis cells. Additionally, a binding partner of MVN was identified in the lipopolysaccharide fraction of M. aeruginosa PCC7806, which involved in the Microcystis colony formation [55]. Together, the toxin-, EPS-, and lectin-related genes may explain the reason why M. panniformis FACHB1757 usually aggregates and produces a larger colony in Lake Taihu during water blooms. In the photo-regulation cluster, psb, apc and gvp with the exception of gvpC were all detected. It is interesting that gvpC is absent from M. panniformis FACHB1757, because this gene encodes GvpC, which is a highly conserved expressed protein in some genera that is closely related to gas vesicles [5658]. Genes related to nutrient assimilation include ntc, pst, and sph clusters. ntcB, pstA, pstB1, pstB2, and pstC were only absent from M. aeruginosa PCC7806 among the four strains, which may be accounted for by the incompleteness of the strain’s genome. Detailed information about function and coordinates of each gene are shown in (Additional file 1: Table S3).

Genome structure and constitution comparison

The genomes of M. aeruginosa NIES843 and M. aeruginosa NIES2549 have no plasmids, whereas a 38 Kb plasmid with a 43.97 % GC content was detected in M. panniformis FACHB1757 in this study. The stable presence of plasmids may play an important role in some Microcystis obtaining competitive advantages [5961]. Microcystis aeruginosa NIES-843 is the first strain of the genus Microcystis to be sequenced for its complete genome with the ABI 3770xl sequencer. Since then, the second completed Microcystis genome (of M. aeruginosa NIES-2549) was released on the April 29, 2015. Thus, the whole genome at the nucleic acid level was compared between M. panniformis FACHB 1757 and M. aeruginosa NIES-843. Mauve, which was designed for identification and alignment of conserved genomic sequences with rearrangements and horizontal transfer, was used to conduct comparative genomic sequence analysis [62]. As shown in (Additional file 1: Figure S1), M. panniformis FACHB1757 underwent extensive chromosome structure rearrangement, which indicates that Microcystis genomes are highly plastic [36].

Self-defense system

Restriction modification system

Comparison with REBASE [63], a restriction enzyme database containing information about restriction enzymes, revealed that DNA methyltransferases and related proteins are involved in the biological process of R–M, and 277 restriction enzymes were found. Detailed classification revealed that 12 and 130 enzymes belonged to type I and type II systems, respectively, which together represented 46.93 % of all enzymes, and are categories of rapidly evolving genes [64]. Sixty-three, 10, and 2 enzymes, respectively, belonged to type IIG, type III, and type IV systems, and one control protein restriction enzyme and 58 unknown enzymes were also found.

Methylation modification analysis

It is widely thought that methylation modification is associated with R-M systems and participates in self-defense against foreign genome invasion. Genome methylation modification and methyl-transferase recognition sequence motifs were analyzed using SMRT (version 2.3.0). In the chromosome, 3204 m4C (N4-methylcytosine), 9,758 m6A (N6-methyladenine), and 31,845 other modified bases were marked as modified (details are available in Additional file 3). Corresponding motif information is included in Table 5.

Table 5 Sequence structure and general information of motifs in the whole genome

CRISPR system

MinCED derived from the CRT [65], was used to predict CRISPR structure. CRISPR are extensively found in prokaryotes and are thought to compose a CRISPR-associated system, which is a putative immune system based on RNA-interference [66]. Three candidate CRISPR clusters on chromosome sequence were annotated under strict parameter and 1 CRISPR on plasmid (further information is available in Additional file 4).

Genomic islands

GEIs are particularly influential in microorganism genomes with regard to virulence, antibiotic resistance, metabolic, symbiosis, or other important adaptations [67]. GEIs have substantial roles in horizontal gene transfer, which is now widely acknowledged as an important force that shapes bacterial genome structure. Island Viewer (version 2.0) [68] was used to predict the GEIs in M. panniformis FACHB1757. Island Viewer integrates SIGI-HMM, Island Pick, and Island Path-DIMOB and built-in databases, including the Virulence Factor Database and Antibiotic Resistance Gene Database. Thirty-six GEIs were found using Island Viewer, and their positions are shown in Fig. 6. Different kinds of functions were identified and are summarized in Table 6. Transposases were identified in most of the GEIs, as they participated in horizontal gene transfer. Toxin-related gene clusters were annotated in six GEIs and probably affect competitiveness and fitness. Some functional genes, such as hat/hatR, were also detected, which indicates the enhanced adaptability and metabolic versatility in this strain.

Fig. 6
figure 6

GEIs distribution in the chromosome of M. panniformis FACHB1757. From inside to outside, green bars illustrate IslandPick prediction, orange bars show the results annotated by SIGI-HMM, and blue bars are predicted by IslandPath-DIMOB. Red bars indicate integrated GEIs candidate positions. Black line plot around the small circle reveal the GC content

Table 6 Functions and types of all 36 GEIs in chromosome

Conclusions

This study presents the complete whole genome sequence of a newly recorded species in China, M. panniformis, and demonstrates several genomic perspectives, including comparison with nine other water bloom-forming cyanobacterial species. A 5.6 Mbp chromosome with a 38 Kbp plasmid was reported, and gene function, methylation modification, CRISPR, and GEIs throughout the genome were described. Large-scale of structure variation was demonstrated by comparison with M. aeruginosa genomes. A Venn diagram of four Microcystis strains showed gene quantity and category variation as a result of evolutionary divergence and revealed that Microcystis has underwent flexible genome evolution.

Abbreviations

CRISPR:

Clustered Regularly Interspaced Short Palindromic Repeats

CRT:

CRISPR Recognition Tool

EPS:

Extracellular polysaccharide

GEIs:

Genome islands

MinCED:

Mining CRISPRs in Environmental Datasets

R–M:

Restriction–modification

TRs:

Tandem repeats

References

  1. Bister B, Keller S, Baumann HI, Nicholson G, Weist S, Jung G, et al. Cyanopeptolin 963A, a Chymotrypsin Inhibitor of Microcystis PCC 7806. J Nat Prod. 2004;67:1755–7. PubMed http://dx.doi.org/10.1021/np049828f.

    Article  PubMed  CAS  Google Scholar 

  2. Kardinaal WEA,Visser PM. Dynamics of caynobacterial toxins. In: Codd G, Lindsay J, Young F, Morrison L, Metcalf J, editors. Harmful Cyanobacteria. Dordrecht, The Netherlands; 2005.p.25-40. http://dx.doi.org/10.1007/1-4020-3022-3_1.

  3. Portmann C, Blom JF, Gademann K, Jüttner F. Aerucyclamides A and B: Isolation and Synthesis of Toxic Ribosomal Heterocyclic Peptides from the Cyanobacterium Microcystis aeruginosa PCC 7806. Journal of Natural Products. 2008;71:1193–1196. PubMed http://dx.doi.org/10.1021/np800118g.

  4. Whitton BA. Microcystis. In: Šejnohová L, Maršálek, editors. Ecology of Cyanobacteria II: Their Diversity in Space and Time. Dordrecht, The Netherlands; 2012.p.195-228. http://dx.doi.org/10.1007/978-94-007-3855-3.

  5. Paerl HW, Huisman J. Climate Blooms like it hot. Science. 2008;320:57–8. PubMed http://dx.doi.org/10.1126/science.1155398.

    Article  PubMed  CAS  Google Scholar 

  6. Weirich CA, Miller TR. Freshwater harmful algal blooms: toxins and children's health. Curr Probl Pediatr Adolesc Health Care. 2014;44:2–24. PubMed http://dx.doi.org/10.1016/j.cppeds.2013.10.007.

    Article  PubMed  Google Scholar 

  7. Trevino-Garrison I, DeMent J, Ahmed FS, Haines-Lieber P, Langer T, Ménager H, et al. Human illnesses and animal deaths associated with freshwater harmful algal blooms-kansas. Toxins. 2015;7:353. PubMed http://dx.doi.org/10.3390/toxins7020353.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  8. Jason S,Hayes PK. Molecular axonomy of Harmful Algae.In: Granéli E, Turner JT. editors. Ecology of Harmful Algae. Dordrecht, The Netherlands; 2006. p. 15–16. http://dx.doi.org/10.1007/978-3-540-32210-8.

  9. Sanchis D, Padilla C, Del Campo FF, Quesada A, Sanz-Alferez S. Phylogenetic and morphological analyses of Microcystis strains (Cyanophyta/Cyanobacteria) from a Spanish water reservoir. NOVA HEDWIGIA. 2005;81:431–48. http://dx.doi.org/10.1127/0029-5035/2005/0081-0431.

    Article  Google Scholar 

  10. Kondo R, Yoshida T, Yuki Y, Hiroishi S. DNA-DNA reassociation among a bloom-forming cyanobacterial genus, Microcystis. Int J Syst Evol Microbiol. 2000;50:767–70. PubMed http://dx.doi.org/10.1099/00207713-50-2-767.

    Article  PubMed  CAS  Google Scholar 

  11. Otsuka S, Suda S, Shibata S, Oyaizu H, Matsumoto S, Watanabe MM. A proposal for the unification of five species of the cyanobacterial genus Microcystis Kutzing ex Lemmermann 1907 under the rules of the Bacteriological Code. Int J Syst Evol Microbiol. 2001;51:873–9. PubMed http://dx.doi.org/10.1099/00207713-51-3-873.

    Article  PubMed  CAS  Google Scholar 

  12. Otsuka S, Suda S, Li R, Watanabe M, Oyaizu H, Matsumoto S, et al. Phylogenetic relationships between toxic and non-toxic strains of the genus Microcystis based on 16S to 23S internal transcribed spacer sequence. FEMS Microbiol Lett. 1999;172:15–21. PubMed http://dx.doi.org/10.1111/j.1574-6968.1999.tb13443.x.

    Article  PubMed  CAS  Google Scholar 

  13. Le Ai NV, Tanabe Y, Matsuura H, Kaya K, Watanabe MM. Morphological, biochemical and phylogenetic assessments of water-bloom-forming tropical morphospecies of Microcystis (Chroococcales, Cyanobacteria). Phycol Res. 2012;60:208–22. http://dx.doi.org/10.1111/j.1440-1835.2012.00650.x.

    Article  Google Scholar 

  14. Bittencourt-Oliveira MC, Moura AN, Gouvêa-Barros S, Pinto E. HIP1 DNA fingerprinting in Microcystis panniformis (Chroococcales, Cyanobacteria). Phycologia. 2007;46:3–9. http://dx.doi.org/10.2216/06-01.1.

    Article  Google Scholar 

  15. Komárek J, Komárková-Legnerova J, Sant’Anna CL, Azevedo MTDP, Senna PAC. Two common Microcystis species (Chroococcales, Cyanobacteria) from tropical America, including M.panniformis sp. nov. Cryptogamie. Algologie. 2002;23:159–77.

    Google Scholar 

  16. White SH, Fabbro LD, Duivenvoorden LJ. Changes in cyanoprokaryote populations, Microcystis morphology, and microcystin concentrations in Lake Elphinstone (Central Queensland, Australia). Environ Toxicol. 2003;18:403–12. PubMed http://dx.doi.org/10.1002/tox.10142.

    Article  PubMed  CAS  Google Scholar 

  17. Sant'Anna CL, Maria Teresa de PA, Pedro Américo CS, Komárek J, Komárková J. Planktic Cyanobacteria from São Paulo State, Brazil: Chroococcales Cianobactérias planctônicas do Estado de São Paulo, Brasil: Chroococcales. Revista Brasileira de Botânica. 2004; 27:213–227. http://dx.doi.org/10.1590/S0100-84042004000200002.

  18. Nguyen LTT, Cronberg G, Larsen J, Moestrup O. Planktic cyanobacteria from freshwater localities in Thuathien-Hue Province, Vietnam I. Morphology and distribution. NOVA HEDWIGIA. 2007;85:1–34. http://dx.doi.org/10.1127/0029-5035/2007/0085-0001.

    Article  Google Scholar 

  19. Borges JT, Sparrapan R, Guimarães JR, Eberlin MN, Augusti R. Chloroform formation by chlorination of aqueous algae suspensions: online monitoring via membrane introduction mass spectrometry. J Brazilian Cheml Soc. 2008;19:950–5. http://dx.doi.org/10.1590/S0103-50532008000500021.

    Article  CAS  Google Scholar 

  20. Luciana Retz de C, Pipole F, Werner VR, Haywood Dail Laughinghouse IV, Antonio Carlos MC, Rangel M, et al. A toxic cyanobacterial bloom in an urban coastal lake, Rio Grande do Sul state, Southern Brazil. Brazilian J Microbiol. 2008;39:761–9. PubMed http://dx.doi.org/10.1590/S1517-838220080004000031.

    Article  Google Scholar 

  21. Bittencourt-Oliveira MC, Cunha MCC, Moura AN. Genetic polymorphism in brazilian Microcystis spp. (Cyanobacteria) toxic and non-toxic through RFLP-PCR of the cpcBA-IGS. Brazilian Archives Biol Technol. 2009;52:901–9. http://dx.doi.org/10.1590/S1516-89132009000400014.

    Article  CAS  Google Scholar 

  22. Caliskan M. Genetic Diversity of Brazilian Cyanobacteria Revealed by Phylogenetic Analysis.In: Bittencourt-Oliveira M, Piccin-Santos V, editors. Genetic Diversity in Microorganisms. 2012.p. 275–290. http://dx.doi.org/10.5772/33137.

  23. Edvani MS, Cícero TSG, Catarina PSR, Silvia MLB. Ocorrência de Cianobactérias no Reservatório de Mundaú, no Estado de Pernambuco, no Período de Janeiro de 2010 a Novembro de 2011 / Occurrence of Cyanobacteria in Mundaú Reservoir, in Pernambuco State During January 2010 to November 2011. Vigilância Sanitária em Debate: Sociedade. 2013;1:35–42. http://dx.doi.org/10.3395/vd.v1i3.40.

    Google Scholar 

  24. Bittencourt-Oliveira MC, Chia MA, de Oliveira HS B, Araujo MKC, Molica RJR, Dias CTS. Allelopathic interactions between microcystin-producing and non-microcystin-producing cyanobacteria and green microalgae: implications for microcystins production. J Appl Phycol. 2015;27:275–84. http://dx.doi.org/10.1007/s10811-014-0326-2.

    Article  CAS  Google Scholar 

  25. Bittencourt-Oliveira MC, Kujbida P, Cardozo KHM, Carvalho VM, Moura AN, Colepicolo P, et al. A novel rhythm of microcystin biosynthesis is described in the cyanobacterium Microcystis panniformis Komárek et al. Biochem Biophys Res Commun. 2005;326:687–94. PubMed http://dx.doi.org/10.1016/j.bbrc.2004.11.091.

    Article  CAS  Google Scholar 

  26. Zhang JY, Zhu BC, Wu ZJ, Xu T, Lu ZH. Microcystis panniformis-A newly recorded species of Microcystis in China. J Lake Sci. 2012;24:647–50.

    Article  Google Scholar 

  27. Paerl HW, Paul VJ. Climate change: links to global expansion of harmful cyanobacteria. Water Res. 2012;46:1349–63. PubMed http://dx.doi.org/10.1016/j.watres.2011.08.002.

    Article  PubMed  CAS  Google Scholar 

  28. Paerl HW. Mitigating harmful cyanobacterial blooms in a human-and climatically-impacted world. Life (Basel, Switzerland). 2014; 4:988–1012. PubMed http://dx.doi.org/10.3390/life4040988.

  29. Qin B, Zhu G, Gao G, Zhang Y, Li W, Paerl HW, et al. A drinking water crisis in Lake Taihu, China: linkage to climatic variability and lake management. Environ Manage. 2010;45:105–12. PubMed http://dx.doi.org/10.1007/s00267-009-9393-6.

    Article  PubMed  Google Scholar 

  30. Qin B, Xu P, Wu Q, Luo L, Zhang Y. Environmental issues of Lake Taihu China. Hydrobiologia. 2007;581:3–14. http://dx.doi.org/10.1007/s10750-006-0521-5.

    Article  CAS  Google Scholar 

  31. Komárek J, Komárková J. Review of the European Microcystis-morphospecies (Cyanoprokaryotes) from nature. Czech Phycology, Olomouc. 2002;2:1–24.

    Google Scholar 

  32. Wood SA, Crowe ALM, Ruck JG, Wear RG. New records of planktonic cyanobacteria in New Zealand freshwaters. New Zealand J Botany. 2005;43:479–92. http://dx.doi.org/10.1080/0028825X.2005.9512969.

    Article  Google Scholar 

  33. Liu Y, Xu Y, Xiao P, Pan Q, Yu G, Li R. Genetic analysis on Dolichospermum (Cyanobacteria; sensu Anabaena) populations based on the culture-independent clone libraries revealed the dominant genotypes existing in Lake Taihu, China. Harmful Algae. 2014;31:76–81. http://dx.doi.org/10.1016/j.hal.2013.09.012.

    Article  Google Scholar 

  34. Li L, Stoeckert Jr CJ, Roos DS. OrthoMCL: identification of ortholog groups for eukaryotic genomes. Genome Res. 2003;13:2178–89. PubMed http://dx.doi.org/10.1101/gr.1224503.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  35. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. MEGA6: Molecular Evolutionary Genetics Analysis version 6.0. Mol Biol Evol. 2013;30:2725–9. PubMed http://dx.doi.org/10.1093/molbev/mst197.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  36. Frangeul L, Quillardet P, Castets AM, Humbert JF, Matthijs HC, Cortez D, et al. Highly plastic genome of Microcystis aeruginosa PCC 7806, a ubiquitous toxic freshwater cyanobacterium. BMC Genom. 2008;9:274. PubMed http://dx.doi.org/10.1186/1471-2164-9-274.

    Article  Google Scholar 

  37. Rantala A, Fewer DP, Hisbergues M, Rouhiainen L, Vaitomaa J, Borner T, et al. Phylogenetic evidence for the early evolution of microcystin synthesis. Proc Natl Acad Sci U S A. 2004;101:568–73. PubMed http://dx.doi.org/10.1073/pnas.0304489101.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  38. Schirrmeister BE, De Vos JM, Antonelli A, Bagheri HC. Evolution of multicellularity coincided with increased diversification of cyanobacteria and the Great Oxidation Event. Proc Natl Acad Sci U S A. 2013;110:1791–6. PubMed http://dx.doi.org/10.1073/pnas.1209927110.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  39. Schirrmeister BE, Antonelli A, Bagheri HC. The origin of multicellularity in cyanobacteria. BMC Evol Biol. 2011;11:45. PubMed http://dx.doi.org/10.1186/1471-2148-11-45.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Palinska KA, Surosz W. Taxonomy of cyanobacteria: a contribution to consensus approach. Hydrobiologia. 2014;740:1–11. http://dx.doi.org/10.1007/s10750-014-1971-9.

    Article  Google Scholar 

  41. Dvořák P, Poulíčková A, Hašler P, Belli M, Casamatta D, Papini A. Species concepts and speciation factors in cyanobacteria, with connection to the problems of diversity and classification. Biodivers Conserv. 2015;24:1–19. http://dx.doi.org/10.1007/s10531-015-0888-6.

    Article  Google Scholar 

  42. Kasai F, Kawachi M, Erata M, Watanabe MM. NIES-Collection, List of Strains: Microalgae and Protozoa. In: Kasai, F., Kawachi, M., Erata, M., Watanabe, M.M., editors. National Institute for Environmental Studies. Japan, Tsukuba; 2004.P. 257.

  43. Chaisson MJ, Tesler G. Mapping single molecule sequencing reads using basic local alignment with successive refinement (BLASR): application and theory. BMC Bioinformatics. 2012;13:238. PubMed http://dx.doi.org/10.1186/1471-2105-13-238.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  44. Miller JR, Delcher AL, Koren S, Venter E, Walenz BP, Brownley AC, et al. Aggressive assembly of pyrosequencing reads with mates. BMC Bioinformatics. 2008;24:2818–24. PubMed http://dx.doi.org/10.1093/bioinformatics/btn548.

    Article  CAS  Google Scholar 

  45. Chin CS, Alexander DH, Marks P, Klammer AA, Drake J, Heiner C, et al. Nonhybrid, finished microbial genome assemblies from long-read SMRT sequencing data. Nat Methods. 2013;10:563–9. PubMed http://dx.doi.org/10.1038/nmeth.2474.

    Article  PubMed  CAS  Google Scholar 

  46. Benson G. Tandem repeats finder: a program to analyze DNA sequences. Nucleic Acids Res. 1999;27:573–80. PubMed http://dx.doi.org/10.1093/nar/27.2.573.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  47. Aziz RK, Bartels D, Best AA, DeJongh M, Disz T, Edwards RA, et al. The RAST Server: rapid annotations using subsystems technology. BMC Genomics. 2008;9:75. PubMed http://dx.doi.org/10.1186/1471-2164-9-75.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Jones P, Binns D, Chang HY, Fraser M, Li W, McAnulla C, et al. InterProScan 5: genome-scale protein function classification. Bioinformatics (Oxford, England). 2014; 30:1236–1240. PubMed http://dx.doi.org/10.1093/bioinformatics/btu031.

  49. Gan N, Xiao Y, Zhu L, Wu Z, Liu J, Hu C, et al. The role of microcystins in maintaining colonies of bloom-forming Microcystis spp. Environ Microbiol. 2012;14:730–42. PubMed http://dx.doi.org/10.1111/j.1462-2920.2011.02624.x.

    Article  PubMed  CAS  Google Scholar 

  50. Tanabe Y, Watanabe MM. Local expansion of a panmictic lineage of water bloom-forming cyanobacterium Microcystis aeruginosa. PloS One. 2011;6:e17085. PubMed http://dx.doi.org/10.1371/journal.pone.0017085.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  51. Zhu W, Dai X, Li M. Relationship between extracellular polysaccharide (EPS) content and colony size of Microcystis is colonial morphology dependent. Biochem Sys Ecol. 2014;55:346–50. http://dx.doi.org/10.1016/j.bse.2014.04.009.

    Article  CAS  Google Scholar 

  52. Zhao L, Lu L, Li M, Xu Z, Zhu W. Effects of Ca and Mg levels on colony formation and EPS content of cultured M. aeruginosa. Procedia Environmental Sciences. 2011; 10, Part B:1452–1458. http://dx.doi.org/10.1016/j.proenv.2011.09.232.

  53. Liu Y, Wang W, Geng L, Chen Y, Yang Z. Polysaccharide content and morphology of Microcystis aeruginosa in response to changes in metabolic carbon flux. Fresenius Environ Bull. 2011;20:1046–50.

    CAS  Google Scholar 

  54. Li M, Zhu W, Gao L, Lu L. Changes in extracellular polysaccharide content and morphology of Microcystis aeruginosa at different specific growth rates. J Appl Phycol. 2013;25:1023–30. http://dx.doi.org/10.1007/s10811-012-9937-7.

    Article  CAS  Google Scholar 

  55. Kehr JC, Zilliges Y, Springer A, Disney MD, Ratner DD, Bouchier C, et al. A mannan binding lectin is involved in cell-cell attachment in a toxic strain of Microcystis aeruginosa. Mol Microbiol. 2006;59:893–906. PubMed http://dx.doi.org/10.1111/j.1365-2958.2005.05001.x.

    Article  PubMed  CAS  Google Scholar 

  56. Hayes PK, Lazarus CM, Bees A, Walker JE, Walsby AE. The protein encoded by gvpC is a minor component of gas vesicles isolated from the cyanobacteria Anabaena flos-aquae and Microcystis sp. Mol Microbiol. 1988;2:545–52. PubMed.

    Article  PubMed  CAS  Google Scholar 

  57. Mlouka A, Comte K, Castets A-M, Bouchier C, Tandeau de Marsac N. The Gas Vesicle Gene Cluster from Microcystis aeruginosa and DNA rearrangements that lead to loss of cell buoyancy. J Bacteriol. 2004;186:2355–65. PubMed.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  58. Dunton PG, Walsby AE. The diameter and critical collapse pressure of gas vesicles in Microcystis are correlated with GvpCs of different length. FEMS Microbiol Lett. 2005;247:37–43. PubMed.

    Article  PubMed  CAS  Google Scholar 

  59. Hauman J. Is a Plasmid(s) Involved in the Toxicity of Microcystis aeruginosa? In: Carmichael WW, editor.The Water Environment. US; 1981. p. 97–102. http://dx.doi.org/10.1007/978-1-4613-3267-1_7.

  60. Schwabe W, Weihe A, Börner T, Henning M, Kohl J-G. Plasmids in toxic and nontoxic strains of the cyanobacterium Microcystis aeruginosa. Curr Microbiol. 1988;17:133–7.

    Article  CAS  Google Scholar 

  61. Bolch CJS, Blackburn SI, Jones GJ, Orr PT, Grewe PM. Plasmid content and distribution in the toxic cyanobacterial genus Microcystis Kützing ex Lemmermann (Cyanobacteria: Chroococcales). Phycologia. 1997;36:6–11. http://dx.doi.org/10.2216/i0031-8884-36-1-6.1.

    Article  Google Scholar 

  62. Darling AC, Mau B, Blattner FR, Perna NT. Mauve: multiple alignment of conserved genomic sequence with rearrangements. Genome Res. 2004;14:1394–403. PubMed http://dx.doi.org/10.1101/gr.2289704.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  63. Roberts RJ, Vincze T, Posfai J, Macelis D. REBASE-a database for DNA restriction and modification: enzymes, genes and genomes. Nucleic Acids Res. 2010;38:D234–6. PubMed http://dx.doi.org/10.1093/nar/gku1046.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  64. Roberts RJ, Vincze T, Posfai J, Macelis D. REBASE-enzymes and genes for DNA restriction and modification. Nucleic Acids Res. 2007;35:D269–270. PubMed http://dx.doi.org/10.1093/nar/gkl891.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  65. Bland C, Ramsey TL, Sabree F, Lowe M, Brown K, Kyrpides NC, et al. CRISPR recognition tool (CRT): a tool for automatic detection of clustered regularly interspaced palindromic repeats. BMC Bioinformatics. 2007;8:209. PubMed http://dx.doi.org/10.1186/1471-2105-8-209.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Haft DH, Selengut J, Mongodin EF, Nelson KE. A guild of 45 CRISPR-associated (Cas) protein families and multiple CRISPR/Cas subtypes exist in prokaryotic genomes. PLoS Comput Biol. 2005;1:e60. PubMed http://dx.doi.org/10.1371/journal.pcbi.0010060.

    Article  PubMed  PubMed Central  Google Scholar 

  67. Dobrindt U, Hochhut B, Hentschel U, Hacker J. Genomic islands in pathogenic and environmental microorganisms. Nat Rev Microbiol. 2004;2:414–24. PubMed http://dx.doi.org/10.1038/nrmicro884.

    Article  PubMed  CAS  Google Scholar 

  68. Langille MG, Brinkman FS. IslandViewer: an integrated interface for computational identification and visualization of genomic islands. Bioinformatics. 2009;25:664–5. PubMed http://dx.doi.org/10.1093/bioinformatics/btp030.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  69. Field D, Garrity G, Gray T, Morrison N, Selengut J, Sterk P, et al. The minimum information about a genome sequence (MIGS) specification. Nat Biotechnol. 2008;26:541–7. PubMed http://dx.doi.org/10.1038/nbt1360.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  70. Woese CR, Kandler O, Wheelis ML. Towards a natural system of organisms: proposal for the domains Archaea, Bacteria, and Eucarya. Proc Natl Acad Sci U S A. 1990;87:4576–9. PubMed http://dx.doi.org/10.1073/pnas.87.12.4576.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  71. McNeill J, Barrie FR, Burdet HM, Demoulin V, Hawksworth DL, Marhold K, et al. International Code of Botanical Nomenclature. Königstein: A.R.G. Ganter; 2006. p. 1.

    Google Scholar 

  72. Woese CR, Stackebrandt E, Macke TJ, Fox GE. A phylogenetic definition of the major eubacterial taxa. Syst Appl Microbiol. 1985;6:143–51. PubMed http://dx.doi.org/10.1016/S0723-2020(85)80047-3.

    Article  PubMed  CAS  Google Scholar 

  73. Stanier RY, Kunisawa R, Mandel M, Cohen-Bazire G. Purification and properties of unicellular blue-green algae (order Chroococcales). Bacteriol Rev. 1971;35:171–205. PubMed.

    PubMed  CAS  PubMed Central  Google Scholar 

  74. Komárek J, Anagnostidis K. Cyanoprokaryota 1. Teil: Chroococcales. In: Ettl H, Gartner G,Heynig H, Molllenhauer D, editors. Süßwasserflora von Mitteleuropa. 1999.p.164-169.

  75. Geitler L. Cyanophyceae. In: Rabenhorst L, von Deutschland KF, editors. Österreich und der Schweiz Leipzig: Akademische Verlagsgesellschaft, vol. 14. 1932. p. 130–59.

    Google Scholar 

  76. Caumette P, Brochier-Armanet C, Normand P. Taxonomy and Phylogeny of Prokaryotes. In: Bertrand J-C, Caumette P, Lebaron P, Matheron R, Normand P, Sime-Ngando T, editors. Environmental Microbiology: Fundamentals and Applications. Netherlands: Springer; 2015. p. 177–9. http://dx.doi.org/10.1007/978-94-017-9118-2_6.

    Google Scholar 

  77. Via-Ordorika L, Fastner J, Kurmayer R, Hisbergues M, Dittmann E, Komarek J, et al. Distribution of microcystin-producing and non-microcystin-producing Microcystis sp. in European freshwater bodies: detection of microcystins and microcystin genes in individual colonies. Syst Appl Microbiol. 2004;27:592–602. PubMed http://dx.doi.org/10.1078/0723202041748163.

  78. Ashburner M, Ball CA, Blake JA, Botstein D, Butler H, Cherry JM, et al. Gene ontology: tool for the unification of biology. The Gene Ontology Consortium. Nature genetics. 2000; 25:25–29.PubMed. http://dx.doi.org/10.1038/75556.

Download references

Acknowledgements

This work was performed under the authority of Wuxi Environmental Monitoring Center, the State Key Laboratory of Bioelectronics (Southeast University), and Institute of Hydrobiology (The Chinese Academy of Sciences). It was partially supported by Major Science and Technology Program for Water Pollution Control and Treatment (2012ZX07101-013-06) to HZ, the National Natural Science Foundation of China (No.61227803), and the NSFC-JST Project (No.41261140337). We thank Ting Song for mapping the collected strain location (Fig. 1a). We gratefully acknowledge the help of Mr. Ge (Director of Wuxi Municipal Environmental Protection Bureau, China) and Mr. Li (Deputy Director of Wuxi Municipal Environmental Protection Bureau, China) in the aspect of administrative support for this project.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Zu-Hong Lu.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

J-YZ and RG contributed equally to this paper. J-YZ carried out microscopy, sample collection and purification, and drafted the manuscript. RG carried out genomic perspective analysis, the genomic evolution analysis, and conceived the manuscript. HL participated in sample culture and genomic DNA extraction. PX provided information about the M. aeruginosa PCC7806 genome. LD and D-MC participated in genome assembly and annotation. H-JZ and B-CZ contributed technical assistance to sample collection and purification. HL, PX and G-LY helped with discussions. Professor Z-HL and R-HL conceived the study, and its design and coordination, and helped draft the manuscript. All authors read and approved the final manuscript.

Additional files

Additional file 1: Tables S1-3 and Figures S1-5.

Table S1. General information regarding phylogenomics and comparative genomic analysis. Table S2. Function annotation assignment from different databases. Table S3. Conserved gene cluster function and position in the genomes of four Microcystis strains. Figure S1. Whole genome comparison with M. aeruginosa NIES843 at the nucleic acid level. Large-scale structural rearrangement was obvious at the nucleic acid level. Blocks on the line for M. panniformis FACHB1757 show that the sequences of the two species were in the same direction, whereas blocks under the line indicated that the sequences ran in the opposite direction. Figure S2. Length distribution of annotated protein-coding genes. The majority of fractions varied in length from 100 to 150. Figure S3. Subsystems and category distributions of genes. Functional annotation by RAST. Figure S4. Histogram demonstrating functional annotation results against GO. Figure S5. Number of each category in function annotation results against KEGG. (DOCX 599 kb)

Additional file 2:

The COG function annotation in each COG cluster. (XLSX 114 kb)

Additional file 3:

N 4 -methylcytosine, N 6 -methyladenine, and other modified bases throughout the whole chromosome. (XLSX 7893 kb)

Additional file 4:

Annotation of three candidate CRISPR clusters on the chromosome sequence and one CRISPR cluster on the plasmid using strict parameters. (XLSX 14 kb)

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, JY., Guan, R., Zhang, HJ. et al. Complete genome sequence and genomic characterization of Microcystis panniformis FACHB 1757 by third-generation sequencing. Stand in Genomic Sci 11, 11 (2016). https://doi.org/10.1186/s40793-016-0130-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40793-016-0130-5

Keywords