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

Genome sequence of Shinella sp. strain DD12, isolated from homogenized guts of starved Daphnia magna

Abstract

Shinella sp. strain DD12, a novel phosphite assimilating bacterium, has been isolated from homogenized guts of 4 days starved zooplankton Daphnia magna. Here we report the draft genome of this bacterium, which comprises 7,677,812 bp and 7505 predicted protein-coding genes.

Introduction

Shinella sp. strain DD12 was isolated from homogenized guts of 4 days starved zooplankton Daphnia magna in the frame of a study, describing the importance, diversity and stability of bacterial communities inside the Daphnia guts. Structural diversity of the bacterial communities were investigated over time, while D. magna were fed with different food sources or were let starve for 4 days, or starved to death [1, 2].

Daphnia spp. are small filter-feeding cladoceran zooplankton organisms which play the role of key members in the freshwater food webs. Heterotrophic bacteria can contribute significantly to the nutrition of Daphnia species [3, 4]. Furthermore, bacteria compared with many algae, are superior competitors for phosphorus and are often characterized by high P:C values [5]. This suggests that bacteria are a rich source of phosphorus for zooplankton [6].

Female D. magna were grown in water from the oligotrophic and low-phosphorus content (below 10 mg.m-3 concentration of total phosphorus in the water column) Lake Constance. Phosphorus in form of phosphate has been identified as the major limiting agent of phytoplankton growth in this lake [7, 8]. Studies performed in the 1990s, after a long period of active care aiming to lower the phosphorus content in the lake water, showed that the primary production of phytoplankton was not influenced substantially after the decrease of the phosphorus content [9]. This phenomenon together with the fact that some bacteria can assimilate reduced inorganic and organic phosphorus compounds (phosphite [+III] and organophosphonates) under phosphate starvation [1019], led us to investigate the newly isolated Shinella sp. strain DD12 in this aspect.

The genus Shinella was established by An et al., in 2006, with Shinella granuli as type species (Ch06T = JCM 13254T ) [20, 21]. It belongs to the family Rhizobiaceae within Alphaproteobacteria and encompasses the following 6 species currently: S. zoogloeoides , S. granuli , S. fusca , S. kummerowiae , S. daejeonensis and S. yambaruensis [2026]. The taxonomic placement of the genus Shinella is shown in Table 1.

Table 1 Classification and general features of Shinella sp. strain DD12

Shinella sp. strain DD12 was chosen for sequencing as it is able to assimilate phosphite under phosphate starvation and use it as single P- source to support its growth. We also focus on the following specific features of this genome - the assimilation of inorganic and organic phosphonates, providing that the organophosphonates are known to serve not only as P-, but as C- and N-sources for different bacteria. This is the first report on a genome sequence of a member of genus Shinella .

Organism information

Classification and features

Shinella sp. DD12 is an aerobic, motile, Gram-negative, non-spore-forming, rod-shaped, hemoheterotroph and psychrotolerant bacterium.

The cells of strain DD12 are short rounded rods with blunt ends and size of 0.6–1 μm in length, and 0.3–0.5 μm in width. Cells are motile via monotrichous flagellum (Fig. 1, Left).

Fig. 1
figure 1

Images of Shinella sp. strain DD12 using scanning (left) electron microscopy and the appearance of colony morphology on solid (middle), and liquid (right) nutrient agar medium

Shinella sp. DD12 forms colonies within 3 to 5 days, when grown on nutrient agar at 18 °C (Fig. 1, Center). Colonies are circular, raised to convex, smooth milky-white in color, slightly opaque with pronounced translucent halo-like edges. In liquid media cells form white fluffy aggregates with finger-like or tree-like morphology (Fig. 1, Right). The strain grows at the temperature range of 10–30 °C. No growth was observed at 37 °C. At 18 °C the strain grows poorly on nutrient broth. At 21 °C it grows with a doubling time of 54–61 h on nutrient broth. By employing a newly developed chemically defined medium (MDS3) with phosphate as the phosphorus source the doubling time was reduced to 32–33 h at 21 °C. The composition of MDS3 medium and the conditions of the tests for phosphite assimilation are available in Additional file 1.

Shinella sp. strain DD12 is positive for catalase, catalase-peroxidase, β-galactosidase and β-glucosidase activity as described for all members of the genus [20, 23]. Strain DD12 can grow oxidatively with the production of acid on different sugars and sugar alcohols. Shinella sp. strain DD12, like other Shinella species except S. fusca , cannot grow on melibiose or starch [20, 2326]. It does not either grow on inulin as is found for S. kummerowiae , whereas there is no data reported for the rest of Shinella strains. Strain DD12 however, shows some specificity in substrate assimilation, as the lack of growth on D-arabinose, while all Shinella strains can grow on this substrate with exception of S. yambaruensis [20, 26]. Analogously, a weak growth on salicin was observed for Shinella sp. strain DD12, where five of the six Shinella strains cannot grow on this substrate. S. granuli growth on salicin remains undetermined [20, 25].

We compared 16S gene sequences of Shinella sp. DD12 with the non-redundant nucleotide collection of NCBI using NCBI MegaBLAST [27, 28]. This comparison revealed that the strain shares 99 % (1445/1453 bp) and 99 % (1438/1446 bp) sequence identity to the 16S rRNA gene sequences of Rhizobium sp. R-24658, and S. zoogloeoides 81 g, respectively. Figure 2 shows the phylogenetic neighborhood of Shinella sp. DD12 in a 16S rRNA sequence based tree of all Shinella type species.

Fig. 2
figure 2

Phylogenetic tree highlighting the position of Shinella sp. strain DD12, based on 16S rRNA gene sequences

The phylogenetic tree was calculated with MEGA5 [29] using the Maximum Likelihood method based on the Jukes Cantor model [30]. Sequences were downloaded from the RDP [31], aligned by CLUSTALW [32] and tested by the bootstrap approach with 1000 resamplings. The length of the tree branches was scaled according the number of substitutions per site (see size bar). Shinella sp. DD12 clustered together with S.granuli Ch06T KCTC12237.

The minimum information about the genome sequence (MIGS) is provided in Table 1, according to MIGS recommendations [33].

Genome sequencing information

Genome project history

This bacterium was selected for sequencing on the basis of its environmental relevance to issues in global P- and N-cycles, and still widely unrecognized reduced P-cycle in nature. Prior to sequencing, Shinella sp. strain DD12 was tested for growth in a newly developed chemically defined liquid medium MDS3 supplemented with 1 mM sodium phosphite as single P-source. The growth and the phosphite assimilation ability of this isolate were confirmed at physiological level (three successive passages in triplicate). The genome project has been deposited in GenBank database (AYLZ00000000) and as an improved high-quality-draft genome sequence in IMG. Genome Sequencing and annotation were done at Göttingen Genomics Laboratory; while cultivation and analysis were performed at the University of Konstanz. The project information and its association with MIGS version 2.0 compliance [33] are presented in Table 2.

Table 2 Project information

Growth conditions and genomic DNA preparation

Shinella sp. DD12 was grown either in nutrient broth or on nutrient agar. The medium was adjusted to pH 7.0 and autoclaved for 25 min at 125 °C. MDS3 medium was used to assay carbohydrate, phosphite and phosphate assimilation by the strain. The chemical composition of the MDS3 medium is given in Additional file 1.

The genomic DNA of the strain was isolated as follows: the cells from 4 ml of a well grown culture in nutrient broth reaching an OD600 of 0.348 ± 0.050 were harvested at 13,000 × g in a benchtop microfuge for 5 min. Cell pellet was suspended in the cell lysis solution of the Purgene Core Kit B (Qiagen, Hilden, Germany). Further, the genomic DNA extraction processed as recommended by the manufacturer. DNA quantity was determined with NanoDrop ND-1000 to ensure that the concentration is greater than 30 ng/μl. One nanogram of the genomic DNA was used for sequencing.

Genome sequencing and assembly

Extracted DNA was used to prepare shotgun libraries for the Genome Analyzer II (Illumina, San Diego, CA, USA). Libraries were prepared according to the manufacturer protocol. Sequencing resulted in 7,118,226 paired-ends Illumina reads of 112 bp and a 72.54-fold coverage. Reads were trimmed using Trimmomatic 0.32 [34] to remove sequences with quality scores lower than 20 (Illumina 1.9 encoding) and remaining adaptor sequences. The initial hybrid de novo assembly employing the SPAdes 2.5 [35] software resulted in 236 contigs larger than 0.5 kb of which 162 were larger than 1 kb including 139 contigs larger than 3 kb. The final assembly had an N50 value of 97,231 bp and an N90 value of 24,331 bp.

Genome annotation

YACOB and GLIMMER [36] software tools were used for automatic gene prediction. RNAmmer [37] and tRNAscan [38] were used for identification of rRNA and tRNA genes, respectively. Functional annotation of the predicted protein-coding genes was carried out with the IMG/ER system [39] and was manually curated by using the Swiss-Prot, TrEMBL, and InterPro databases [40].

Genome properties

The genome statistics are provided in Table 3. The pseudogenes may also be counted as protein coding or RNA genes, so they are not additive under total gene count.

Table 3 Genome statistics

The draft genome of Shinella sp. DD12 consists of 236 contigs comprising 7.678 Mb and an overall GC content of 63.40 mol%. The genome harbors 7555 putative genes, of which 7505 are protein-encoding and 50 RNAs (2 rRNA and 48 tRNA). The tRNAs included tRNA necessary for selenocystein incorporation (SHLA_2c001070). Protein encoding genes with a putative function prediction are 6241 (82.61 %) of all proteins in the genome and 1264 (16.73 %) were annotated as hypothetical proteins. The majority of the protein-encoding genes 5394 (71.40 %) were assigned to one of the known COG categories [41]. The distribution of these genes with respect to assigned functions is presented in Table 4.

Table 4 Number of genes associated with general COG functional categories

Insights from the genome sequence

The genome of Shinella sp. DD12 consists of a circular chromosome and at least 7 plasmids as we could detect 7 different repABC gene clusters located on 7 different contigs. Further database analysis revealed that all complete sequenced Rhizobiaceae genomes harbor usually between 2 and 6 plasmids, but species with up to 9 plasmids have been found, as in Ensifer fredii HH103 [42].

The strain is aerobe and its aerobic respiratory chain contains all genes encoding Complex I to Complex V. In addition, strain DD12 possesses a complete denitrification pathway via periplasmic cytochrome c [43]. The pathway found in this genome includes the genes encoding a periplasmic nitrate reductase napABC (SHLA_29c000730 - SHLA_29c000770), NO-forming nitrtite reductase nirK (SHLA_5c000410), nitric oxide reductase norCBD (SHLA_5c000290 - SHLA_5c000340) and a nitrous oxide reductase nosZ (SHLA_36c000580). The genome analysis of the strain DD12 revealed the potential abilities of this isolate to reduce nitrogen via the dissimilatory nitrate reduction to ammonia (DNRA) pathway, and to assimilate nitrate to L-glutamine and L-glutamate. The genes encoding nitrogen fixation ability such as N-acetylglucosaminyl transferase (nodC) or nitrogenase reductase (nifH) are absent from the genome. This is consistant with the previously reported lack of nitrogen fixation ability in free-living Shinella species, except for the only known symbiont S. kummerowiae [20, 23, 25].

Shinella sp. strain DD12 is able to utilize reduced inorganic phosphonate (phosphite) and presumably organophosphonates as the single P-sources to support its growth. The phosphite oxidation most probably proceeds through a periplasmic alkaline phosphatase (phoA), analogously to E.coli [18], or through the carbon-phosphorus (C-P) lyase complex [1012, 15]. The latter complex is known to have broader substrate specificity, including the oxidation of phosphite and the assimilation of the most common organophosphonate - methylphosphonate. The C-P lyase complex, although the presence of the conserved structural phnGHIJKLM gene cluster, shows low conservation level of the gene sequences arrangement amongst representatives of Alpha- Beta - Gamma- and Deltaproteobacteria (Fig. 3). However, this drastically changes within the Rhizobiaceae members harboring a C-P lyase complex. The C-P complex shows highest conservation level of the gene sequences and their arrangement amongst the Rhizobiaceae members that harbors it.

Fig. 3
figure 3

Tblastx comparison of the C-P lyase complex. An E-value cutoff of 1e-10 was used and visualization was done with the program Easyfig [55]. Functional genes of the C-P lyase complex were marked in green, accessory genes in orange, the ABC-type transporter in blue and the regulatory subunit in red. Genes not directly associated with this pathway are marked in grey

In addition, the Shinella sp. DD12 genome harbors a 2-aminoethylphosphonate (2-AEP) degradation pathway, which operates through the phosphonoacetaldehyde dehydrogenase - phosphonoacetate hydrolase (phnWAY) [16, 44]. The 2-AEP (ciliatine) is a common phosphonate constituent of the phospholipids in a variety of marine invertebrates, including ciliated protozoa, sees anemones, some plants and animals. Recently, the synthesis of sphingophosphonolipids was found in some bacterial species including Bacteriovorax stolpii , a facultative predator which parasitizes larger Gram-negative bacteria [45]. A Tblastx comparison of the phnWAY encoding operon from Shinella sp. strain DD12 with another 3 species belonging to Alphaproteobacteria , two of which members of Rhizobiaceae is shown on Fig. 4. An analysis of all genomes available at IMG (as of April 1, 2015) against phosphonoacetaldehyde dehydrogenase encoding gene (phnY) revealed its presence in 431 gene clusters. However, the complete phnWAY operon was present in only 92 genomes of which 41 belong to Rhizobiaceae species. Furthermore, the phnWAY operon was placed in the majority of the Alphaproteobacteria genomes in close proximity to the fbpABC transporter involved in the utilization of xenosiderophores as iron sources in a TonB-independent manner. It is known that the fbpABC gene cluster is transcribed as separate operon in Neisseria meningitidis [46]. However, whether this cluster plays a role in phosphonate uptake in the cell is unclear.

Fig. 4
figure 4

Tblastx comparison of Alphaproteobacteria phnWAY operon. An E-value cutoff of 1e-10 was used and visualization was done with the program Easyfig [55]. The phnWAY cluster was marked in orange tones; sodium/phosphonate symporter (yjbB) was marked in purple; the genes coding ABC-type ferric uptake system (fbpABC) were marked in blue; the transcriptional regulator (lysR) is shown in green

Conclusions

The draft genome sequence of Shinella sp. strain DD12 described here is the first genome sequence of a member of the genus Shinella . The genome of the strain DD12 suggests the presence of 7 plasmids, which is often found amongst members of Rhizobiaceae .

The genome analysis of Shinella sp. strain DD12 indicates that the bacterium is a denitrifier, as it harbours two complete sets of genes encoding: i) the dissimilatory nitrate reduction to ammonia pathway and ii) assimilative nitrate reduction to L-glutamine, and L-glutamate pathway. Shinella sp. strain DD12 cannot fix nitrogen, similarly to the other free-living known Shinella species, whereas the symbiotically growing S. kummerowiae is a nitrogen fixing bacterium.

Finally, the genome of Shinella sp. DD12 encodes three complete pathways for assimilation of phosphonates. The presence of these three pathways indicates relatively broad abilities to utilise reduced phosphonates as P- and/or C- and N-sources, compared to the remaining genomes of Rhizobiaceae members and even to Alphaproteobacteria as a whole. This could be a great advantage for the strain DD12 in environments where other bacteria can face growth limitations, providing that the inorganic- and organophosphonates are naturally occurring compounds. Furthermore, the presence of the genes encoding the complete pathway for 2-AEP containing biomolecules might provide a defence mechanism against predator and parasite bacteria.

Abbreviations

RDP:

Ribosomal Database Project (East Lansing, MI, USA)

References

  1. Freese HM, Schink B. Composition and stability of the microbial community inside the digestive tract of the aquatic crustacean Daphnia magna. Microb Ecol. 2011;62(4):882–94. http://dx.doi.org/10.1007/s00248-011-9886-8.

    Article  PubMed  Google Scholar 

  2. Martin-Creuzburg D, Beck B, Freese HM. Food quality of heterotrophic bacteria for Daphnia magna: evidence for a limitation by sterols. FEMS Microbiol Ecol. 2011;76(3):592–601. http://dx.doi.org/10.1111/j.1574-6941.2011.01076.x.

    Article  PubMed  Google Scholar 

  3. Taipale S, Kankaala P, Tiirola M, Jones RI. Whole-lake dissolved inorganic C-13 additions reveal seasonal shifts in zooplankton diet. Ecology. 2008;89(2):463–74. http://www.esajournals.org/doi/abs/10.1890/07-0702.1.

    Article  PubMed  Google Scholar 

  4. Taipale S, Kankaala P, Hamalainen H, Jones RI. Seasonal shifts in the diet of lake zooplankton revealed by phospholipid fatty acid analysis. Freshwater Biol. 2009;54(1):90–104. http://dx.doi.org/10.1111/j.1365-2427.2008.02094.x.

    Article  CAS  Google Scholar 

  5. Vadstein O. Heterotrophic, planktonic bacteria and cycling of phosphorus - Phosphorus requirements, competitive ability, and food web interactions. In: Shink B, editor. Adv Microb Ecol, vol 16. Springer US; 2000. p 115–67.

  6. Hessen DO, Andersen T. Bacteria as a source of phosphorus for zooplankton. Hydrobiologia. 1990;206(3):217–23. http://dx.doi.org/10.1007/BF00014087.

    Article  CAS  Google Scholar 

  7. McDowell MM, Ivey MM, Lee ME, et al. Detection of hypophosphite, phosphite, and orthophosphate in natural geothermal water by ion chromatography. J Chromatogr A. 2004;1039(1–2):105–11. http://www.sciencedirect.com/science/article/pii/S0021967303021368.

    Article  CAS  PubMed  Google Scholar 

  8. Gerhardt S, Boos K, Schink B. Uptake and release of phosphate by littoral sediment of a freshwater lake under the influence of light or mechanical perturbation. 2010;69(1). http://www.jlimnol.it/index.php/jlimnol/article/view/jlimnol.2010.54.

  9. Tilzer MM, Gaedke U, Schweizer A, Beese B, Wieser T. Interannual variability of phytoplankton productivity and related parameters in Lake Constance: no response to decreased phosphorus loading? J Plankton Res. 1991;13(4):755–77. http://plankt.oxfordjournals.org/content/13/4/755.

    Article  Google Scholar 

  10. Kokonova SV, Nesmeyanova MA. Phosphonates and their degradation by microorganisms. Biochemistry (Moscow). 2002;67(2):184–95. http://dx.doi.org/10.1023/A%3A1014409929875.

    Article  Google Scholar 

  11. White AK, Metcalf WW. Microbial metabolism of reduced phosphorus compounds. Annu Rev Microbiol. 2007;61:379–400.

    Article  CAS  PubMed  Google Scholar 

  12. Metcalf WW, Wanner BL. Involvement of the Escherichia coli phn (psiD) gene cluster in assimilation of phosphorus in the form of phosphonates, phosphite, Pi esters, and Pi. J Bacteriol. 1991;173(2):587–600. http://jb.asm.org/cgi/content/abstract/173/2/587.

    PubMed Central  CAS  PubMed  Google Scholar 

  13. Simeonova DD, Wilson MM, Metcalf WW, Schink B. Identification and heterologous expression of genes involved in anaerobic dissimilatory phosphite oxidation by Desulfotignum phosphitoxidans. J Bacteriol. 2010;192(19):5237–44. http://jb.asm.org/cgi/content/abstract/192/19/5237.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  14. Poehlein A, Daniel R, Schink B, Simeonova DD. Life based on phosphite: a genome-guided analysis of Desulfotignum phosphitoxidans. BMC Genomics. 2013;14(1):753. http://www.biomedcentral.com/1471-2164/14/753.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  15. Villarreal-Chiu JF, Quinn JP, McGrath JW. The genes and enzymes of phosphonate metabolism by bacteria, and their distribution in the marine environment. Front Microbiol. 2012;3. http://www.frontiersin.org/Journal/Abstract.aspx?s=53&name=aquatic_microbiology&ART_DOI=10.3389/fmicb.2012.00019.

  16. Borisova SA, Christman HD, Metcalf MEM, Zulkepli NA, Zhang JK, van der Donk WA, et al. Genetic and biochemical characterization of a pathway for the degradation of 2-aminoethylphosphonate in Sinorhizobium meliloti 1021. J Biol Chem. 2011;286(25):22283–90. http://www.jbc.org/content/286/25/22283.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  17. Dumora C, Lacoste A-M, Cassaigne A. Phosphonoacetaldehyde hydrolase from Pseudomonas aeruginosa: Purification properties and comparison with Bacillus cereus enzyme. Biochim Biophys Acta. 1989;997(3):193–8. http://www.sciencedirect.com/science/article/pii/0167483889901866.

    Article  CAS  PubMed  Google Scholar 

  18. Yang K, Metcalf WW. A new activity for an old enzyme: Escherichia coli bacterial alkaline phosphatase is a phosphite-dependent hydrogenase. Proc Natl Acad Sci U S A. 2004;101(21):7919–24. http://www.pnas.org/content/101/21/7919.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  19. Poehlein A, Freese HM, Daniel R, Simeonova DD. Draft Genome Sequence of Serratia sp. Strain DD3, Isolated from the Guts of Daphnia magna. Genome Announc. 2014;2(5). http://genomea.asm.org/content/2/5/e00903-14.abstract.

  20. An D-S, Im W-T, Yang H-C, Lee S-T. Shinella granuli gen. nov., sp. nov., and proposal of the reclassification of Zoogloea ramigera ATCC 19623 as Shinella zoogloeoides sp. nov. Int J Syst Evol Microbiol. 2006;56(2):443–8. http://ijs.sgmjournals.org/content/56/2/443.

    Article  CAS  PubMed  Google Scholar 

  21. Carrareto Alves LM, Marcondes de Souze JA, de Mello Varani A, de Macedo Lemos EG. The Family Rhizobiaceae. In: Rosenberg E, DeLong EF, Lory S, Stackebrandt E, Thompson F, editors. Berlin Heidelberg: Springer-Verlag; 2014.

  22. Euzéby JP. List of bacterial names with standing in nomenclature: a folder available on the Internet. Int J Syst Evol Microbiol 1997, 47(2):590-592. http://ijs.microbiologyresearch.org/content/journal/ijsem/10.1099/00207713-47-2-590. URL: http://www.bacterio.net.

  23. Lin DX, Wang ET, Tang H, Han TX, He YR, Guan SH, et al. Shinella kummerowiae sp. nov., a symbiotic bacterium isolated from root nodules of the herbal legume Kummerowia stipulacea. Int J Syst Evol Microbiol. 2008;58(6):1409–13. http://ijs.sgmjournals.org/content/58/6/1409.abstract.

    Article  CAS  PubMed  Google Scholar 

  24. Lee M, Woo S-G, Ten LN. Shinella daejeonensis sp. nov., a nitrate-reducing bacterium isolated from sludge of a leachate treatment plant. Int J Syst Evol Microbiol. 2011;61(9):2123–8. http://ijs.sgmjournals.org/content/61/9/2123.abstract.

    Article  CAS  PubMed  Google Scholar 

  25. Vaz-Moreira I, Faria C, Lopes AR, Svensson LA, Moore ERB, Nunes OC, et al. Shinella fusca sp. nov., isolated from domestic waste compost. Int J Syst Evol Microbiol. 2010;60(1):144–8. http://ijs.sgmjournals.org/content/60/1/144.abstract.

    Article  CAS  PubMed  Google Scholar 

  26. Matsui T, Shinzato N, Tamaki H, Muramatsu M, Hanada S. Shinella yambaruensis sp. nov., a 3-methyl-sulfolane-assimilating bacterium isolated from soil. Int J Syst Evol Microbiol. 2009;59(3):536–9. http://ijs.sgmjournals.org/content/59/3/536.abstract.

    Article  CAS  PubMed  Google Scholar 

  27. Zhang Z, Schwartz S, Wagner L, Miller W. A greedy algorithm for aligning DNA sequences. J Comput Biol. 2000;7(1-2):203–2014. http://online.liebertpub.com/doi/abs/10.1089/10665270050081478.

    Article  CAS  PubMed  Google Scholar 

  28. Morgulis A, Coulouris G, Raytselis Y, Madden TL, Agarwala R, Schäffer AA. Database indexing for production MegaBLAST searches. Bioinformatics. 2008;24(16):1757–64. http://bioinformatics.oxfordjournals.org/content/24/16/1757.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  29. Tamura K, Peterson D, Peterson N, Stecher G, Nei M, Kumar S. MEGA5: Molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol Biol Evol. 2011;28(10):2731–9. http://mbe.oxfordjournals.org/content/28/10/2731.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  30. Jukes TH, Cantor CR. Evolution of protein molecules. In: Munro HN, editor. Mammalian Protein Metabolism. edition. New York: Academic; 1969. p. 21–132.

    Chapter  Google Scholar 

  31. Cole JR, Chai B, Farris RJ, Wang Q, Kulam SA, McGarrell, DM, et al. The Ribosomal Database Project (RDP-II): sequences and tools for high-throughput rRNA analysis. Nucleic Acids Res. 2005;33 suppl 1:D294–6. http://nar.oxfordjournals.org/content/33/suppl_1/D294.

    PubMed Central  CAS  PubMed  Google Scholar 

  32. Larkin MA, Blackshields G, Brown NP, Chenna R., McGettigan PA, McWilliam H, et al. Clustal W and Clustal X version 2.0. Bioinformatics. 2007;23(21):2947–8. http://bioinformatics.oxfordjournals.org/content/23/21/2947.

    Article  CAS  PubMed  Google Scholar 

  33. 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(5):541–7. http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=18464787.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  34. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30(15):2114–20. http://bioinformatics.oxfordjournals.org/content/30/15/2114.abstract.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  35. Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, et al. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J Comput Biol. 2012;19(5):455–77. http://online.liebertpub.com/doi/abs/10.1089/cmb.2012.0021.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  36. Tech M, Merkl R. YACOP: enhanced gene prediction obtained by a combination of existing methods. In Silico Biol. 2003;3(4):441–51. http://iospress.metapress.com/content/EHTJD53AEG66U72E.

    CAS  PubMed  Google Scholar 

  37. Lagesen K, Hallin P, Rødland EA, Stærfeldt H-H, Rognes T, Ussery DW. RNAmmer: consistent and rapid annotation of ribosomal RNA genes. Nucleic Acids Res. 2007;35(9):3100–8. http://nar.oxfordjournals.org/content/35/9/3100.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  38. Lowe TM, Eddy SR. tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res. 1997;25(5):0955–964. http://nar.oxfordjournals.org/content/25/5/0955.

    Article  CAS  Google Scholar 

  39. Markowitz VM, Mavromatis K, Ivanova NN, Chen I-MA, Chu K, Kyrpides NC. IMG ER: a system for microbial genome annotation expert review and curation. Bioinformatics. 2009;25(17):2271–8. http://bioinformatics.oxfordjournals.org/content/25/17/2271.

    Article  CAS  PubMed  Google Scholar 

  40. Zdobnov EM, Apweiler R. InterProScan – an integration platform for the signature-recognition methods in InterPro. Bioinformatics. 2001;17(9):847–8. http://bioinformatics.oxfordjournals.org/content/17/9/847.abstract.

    Article  CAS  PubMed  Google Scholar 

  41. Tatusov RL, Koonin EV, Lipman DJ. A genomic perspective on protein families. Science. 1997;278(5338):631–7. http://www.sciencemag.org/content/278/5338/631.

    Article  CAS  PubMed  Google Scholar 

  42. Weidner S, Becker A, Bonilla I, Jaenicke S, Lloret J, Margaret I, et al. Genome sequence of the soybean symbiont Sinorhizobium fredii HH103. J Bacteriol. 2012;194(6):1617–8. http://jb.asm.org/content/194/6/1617.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  43. Shin YK, Hirashi A, Sugiyama J. Molecular systematics of the genus Zoogloea and emendation of the genus. Int J Syst Bacteriol. 1993;43(4):826–31. http://ijs.sgmjournals.org/content/43/4/826.

    Article  CAS  PubMed  Google Scholar 

  44. Jiang W, Metcalf WW, Lee KS, Wanner BL. Molecular cloning, mapping, and regulation of Pho regulon genes for phosphonate breakdown by the phosphonatase pathway of Salmonella typhimurium LT2. J Bacteriol. 1995;177(22):6411–21. http://jb.asm.org/content/177/22/6411.

    PubMed Central  CAS  PubMed  Google Scholar 

  45. Watanabe Y, Nakajima M, Hoshino T, Jayasimhulu K, Brooks E, Kaneshiro E. A novel sphingophosphonolipid head group 1-hydroxy-2-aminoethyl phosphonate in Bdellovibrio stolpii. Lipids. 2001;36(5):513–9. http://dx.doi.org/10.1007/s11745-001-0751-3.

    Article  CAS  PubMed  Google Scholar 

  46. Khun HH, Deved V, Wong H, Lee BC. fbpABC gene cluster in Neisseria meningitidis is transcribed as an operon. Infect Immun. 2000;68(12):7166–71. http://www.ncbi.nlm.nih.gov/pmc/articles/PMC97834/.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  47. Woese CR, Kandler O, Wheelis ML. Towards a natural system of organisms: proposal for the domains Archaea, Bacteria, and Eucarya. Proc Natl Acad Sci. 1990;87(12):4576–9. http://www.pnas.org/content/87/12/4576.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  48. Garrity GM, Bell JA, Lilburn T. Phylum XIV. Proteobacteria phyl. nov. In: Garrity GM, Brenner DJ, Krieg NR, Staley JT, editors. Bergey's Manual of Systematic Bacteriology, vol. 2. 2nd ed. New York: Springer; 2005. p. 1.

    Chapter  Google Scholar 

  49. Garrity GM, Bell JA, Lilburn T. Class I. Alphaproteobacteria class.nov. In: Garrity GM, Brenner DJ, Krieg NR, Staley JT, editors. Bergey's Manual of Systematic Bacteriology, vol. 2. 2nd ed. New York: Springer; 2005. p. 1.

    Chapter  Google Scholar 

  50. Kuykendall LD. Order VI. Rhizobiales ord. nov. In: Garrity GM, Brenner DJ, Krieg NR, Staley JT, editors. Bergey's Manual of Systematic Bacteriology, vol. 2. 2nd ed. New York: Springer; 2005. p. 324.

    Google Scholar 

  51. Skerman VBD, McGowan V, Sneath PHA. Approved lists of bacterial names. Int J Syst Bacteriol. 1980;30(1):225–420. http://ijs.sgmjournals.org/content/30/1/225.

    Article  Google Scholar 

  52. Conn HJ. Taxonomic relationships of certain non-sporeforming rods in soil. In: Agricultural and inductrial bacteriology. J Bacteriol. 1938;36(3):303–24. http://jb.asm.org/content/36/3/303.

    Google Scholar 

  53. Xie C-H, Yokota A. Zoogloea oryzae sp. nov., a nitrogen-fixing bacterium isolated from rice paddy soil, and reclassification of the strain ATCC 19623 as Crabtreella saccharophila gen. nov., sp. nov. Int J Syst Evol Microbiol. 2006;56(3):619–24. http://ijs.sgmjournals.org/content/56/3/619.abstract.

    Article  CAS  PubMed  Google Scholar 

  54. Ashburner M, Ball CA, Blake JA, Botstein D, Butler H, Cherry JM. Gene ontology: tool for the unification of biology. The Gene Ontology Consortium. Nat Genet. 2000;25(1):25–9. http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Citation&list_uids=10802651.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  55. Sullivan MJ, Petty NK, Beatson SA. Easyfig: a genome comparison visualizer. Bioinformatics. 2011;27(7):1009–10. http://bioinformatics.oxfordjournals.org/content/27/7/1009.abstract.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank Heike M. Freese, for supplying the strain, Bernhard Schink for support, Joachim Hentschel for the scaning electron micrograph of the isolate, Bernd Gahr, Sylke Wiechmann, Frauke-Dorothee Meyer, Gabriele Pötter and Kathleen Gollnow for technical assistance. This work was supported by the University of Konstanz. D.D.S. was supported partially by a grant of the Deutsche Forschungsgemeinschaft, Bonn – Bad Godesberg, Germany (SI 1300/4-1). We acknowledge the support of the Open Publication Funds of the University of Konstanz.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Diliana D. Simeonova.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

DDS and AP designed research. DDS characterized the strain DD12. HF has contributed for the chemotaxonomy. AP carried out genome analyses with the help of DDS. DDS, AP and RD wrote the manuscript. All authors read and approved the final manuscript.

Additional file

Additional file 1:

Composition of MDS3 medium and growth conditions in phosphite assimilation tests. (PDF 64.3 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

Poehlein, A., Freese, H., Daniel, R. et al. Genome sequence of Shinella sp. strain DD12, isolated from homogenized guts of starved Daphnia magna . Stand in Genomic Sci 11, 14 (2016). https://doi.org/10.1186/s40793-015-0129-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40793-015-0129-3

Keywords