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

Insights into the single cell draft genome of “Candidatus Achromatium palustre”

Abstract

Candidatus Achromatium palustre” was recently described as the first marine representative of the Achromatium spp. in the Thiotrichaceae - a sister lineage to the Chromatiaceae in the Gammaproteobacteria. Achromatium spp. belong to the group of large sulfur bacteria as they can grow to nearly 100 μm in size and store elemental sulfur (S0) intracellularly. As a unique feature, Achromatium spp. can accumulate colloidal calcite (CaCO3) inclusions in great amounts. Currently, both process and function of calcite accumulation in bacteria is unknown, and all Achromatium spp. are uncultured. Recently, three single-cell draft genomes of Achromatium spp. from a brackish mineral spring were published, and here we present the first draft genome of a single “Candidatus Achromatium palustre” cell collected in the sediments of the Sippewissett Salt Marsh, Cape Cod, MA. Our draft dataset consists of 3.6 Mbp, has a G + C content of 38.1 % and is nearly complete (83 %). The next closest relative to the Achromatium spp. genomes is Thiorhodovibrio sp. 907 of the family Chromatiaceae, containing phototrophic sulfide-oxidizing bacteria.

Introduction

Achromatium spp. have been known for over a century and have been detected in sediments of freshwater [15] and marine [6, 7] environments. They are large rod-shaped bacteria that typically range in size from 5–40 μm in diameter and 15–100 μm in length, and they migrate by slow rolling along the opposing sedimentary redox gradients of sulfide and oxygen [8]. The first species described was Achromatium oxaliferum , named after the large intracellular inclusions, which were suggested to consist of calcium oxalate [5]. Later it was found that they are actually composed of calcium carbonate, also referred to as calcite [1, 3, 9]. To this day, Achromatium spp. remain uncultured and their ecophysiology has been investigated in freshwater populations, mainly using microcosm experiments [2, 8, 1013]. Achromatium spp. are presumably chemolithotrophic, and oxidize reduced sulfur compounds completely to sulfate [11, 13, 14], they are suggested to be microaerophilic, and may use nitrate as alternative electron acceptor to oxygen [3, 10, 1316].

A marine population of Achromatium spp. [6] was recently described in more detail [7] and this population showed altered migration patterns as well as an increased tolerance to oxygen as reported for freshwater populations [14]. Besides calcite and sulfur inclusions, staining and energy dispersive X-ray analysis revealed a third type of inclusion in the salt marsh Achromatium containing a high concentration of Ca2+ ions that were suggested to be stored for the rapid, dynamic precipitation of calcium carbonate. The number of inclusions varied according to the position of a cell relative to the redox gradient of the sediment [7].

Sequencing Achromatium genomes not only provides insight into the genetic and ecophysiological potential of these uncultured organisms in order to find genetic evidence supporting field and microcosm observations (Table 1), but also enables the identification of candidate genes involved in calcite accumulation. Three draft genomes of Achromatium from a mineral spring in Florida were recently published [17], and here we present the first draft genome of a marine Achromatium representative.

Table 1 Classification and general features of “Candidatus Achromatium plaustre” according to the MIGS recommendations [40]

Organism information

Classification and features

As the most striking phenotypic feature, Candidatus A. palustre, as well as other described Achromatium species, appear bright white to the naked eye, as they contain multiple intracellular calcium carbonate (CaCO3) inclusions, and elemental sulfur (S0) granules, that fill nearly the entire interior of the cell. There is no large central vacuole as observed in other large sulfur bacteria, e.g. Beggiatoa spp. [18]. Calcite inclusions vary in diameter, but are typically several micrometers in size. Under the microscope, Achromatium spp. appear bulgy and rock-like (Fig. 1a), and one can observe the slowly jerky rolling motility of the large cells. TEM imaging of freshwater Achromatium showed that the calcite inclusions have a central nucleation point that is surrounded by concentric rings of precipitated calcite, and that they are probably enclosed by a membrane [14]. The salt marsh Achromatium were on average 20 × 26 μm in diameter, rod-shaped, contained several large calcite inclusions, and numerous small interstitial inclusions. Some cells had an external sheath, supposedly a layer of mucus, to which occasionally other rod-shaped and filamentous bacteria were attached [7]. Staining with Calcium Orange-5 N (Fig. 1c), or Calcium Green-1 revealed additional inclusions that were highly enriched in Ca2+ and of much smaller size (<1 μm) in the interstitial space between the large calcite inclusions (compare Fig. 1b and c) [7]. Achromatium have a Gram-negative cell wall [3, 19], and the cytoplasm as well as DNA is distributed across the entire cell in thin (<2 μm) threads stretching between the inclusions [7].

Fig. 1
figure 1

Micrographs of Candidatus Achromatium palustre. a Light micrograph showing that each cell contains large bulgy calcite inclusions, which highly reflect the light. The square-shaped, reflective organism in the top middle is a diatom. b and c show the same cell imaged with a confocal microscope; b is taken with transmitted light showing smaller inclusions between the large calcite inclusions, and c is the fluorescent signal of Calcium Orange-5 N showing the co-localization of highly concentrated Ca2+ ions (bright red) with the smaller granules visible in (b)

Candidatus Achromatium palustre was detected in Little Sippewissett Salt Marsh on Cape Cod, Massachusetts, where they occurred mainly in the upper 2 cm of the sediment of a tide pool. From the deeper layers of the flocculous, organic-rich phytodetritus, high sulfide concentrations diffused upwards meeting the sediment/water interface during the night. During the day, photosynthetic algae and cyanobacteria generated supersaturated oxygen concentrations in the surficial sediment and overlying water column, which created an oxic, sulfide-free zone in the upper millimeters of the sediment [7].

The salt marsh Achromatium population co-occurred with highly abundant and conspicuous, millimeter-size aggregates of purple sulfur bacteria in the surficial sediment layers. Other phototrophic bacteria (phylum Cyanobacteria ) and eukaryotes (diatoms) are also found in higher densities at the sediment/water interface; heterotrophic sulfate-reducing bacteria of the Deltaproteobacteria dominate in deeper sediment layers [7, 20, 21]. The single Candidatus A. palustre cell was isolated by an initial sieving of the sediment to remove the large aggregates and debris, followed by manual removal of the cell using a glass Pasteur pipette, and a successive washing steps in sterile water until contaminants were out-diluted.

Currently, Achromatium spp. 16S rRNA gene sequences are either classified as Achromatium oxaliferum , or Achromatium sp., intermixed [2, 3, 22] between the two phylogenetic subclusters “A” and “B” (Fig. 2). These subclusters not only separate by 16S rRNA gene sequence difference, but also by the presence (A) or absence (B) of helix 38 in the V6 region [2]. Recently, it was proposed that the subclusters may represent and/or include several candidatus taxa [8], however, due to the lack of cultures, a reclassification of the members of the Achromatium lineage is challenging, as it cannot be based on sequence information alone [23]. With the accumulation of information about the natural populations and subpopulations through culture-independent techniques the phylotypes will most likely receive phylogenetic attention in the future. One subcluster in “cluster B” was already classified as “Candidatus Achromatium minus” based on sequence divergence and morphological difference [24]. “Candidatus Achromatium palustre” was likewise classified as part of “cluster A”, based on 16S rRNA gene sequence information and their adaptation to the very different habitat, as well as their altered behavioural characteristics [7] (Fig. 2).

Fig. 2
figure 2

Phylogenetic tree based on 16S rRNA gene sequence information. The reconstruction was performed originally with 80 sequences, of which only a subset is shown here, and a total of aligned 1,101 positions using the maximum likelihood RaxML method of the ARB software package [49]. The tree was rooted with representatives of the Deltaproteobacteria. Branching patterns supported by <40 % confidence in 100 bootstraps replicates were manually converted into multifurcations. Candidatus Achromatium palustre, the source organism of the here presented genome, affiliates with cluster A in the Achromatium lineage, and is highlighted in bold face. (T) marks type strains/sequences, and asterisks (*) shows the availability of a genome

Achromatium spp. have originally been classified in the family Achromatiaceae [25, 26] as a sister family of the Beggiatoaceae [27] and Leucotrichaceae [28] within the order Beggiatoales [29, 30]. Recently, a reclassification was published [31], merging these families into one newly created family Thiotrichaceae (Table 1), in the order Thiotrichales [32].

Genome sequencing information

Genome project history

The sequencing project was initiated in August 2013, when cells were collected from the field, isolated, and subjected to multiple displacement amplification. The amplified DNA was sequenced in November 2014, the raw data were integrated into the JGI pipeline Jigsaw2.4.1, where they were quality-checked and assembled. Annotation and further decontamination was performed through IMG [33]. After final analysis for contamination and completion in CheckM [34], the draft genome (Table 2) was completed in February 2015, when it was deposited in the Genome On-Line Database and became available in IMG (Ga0065144). The whole genome shotgun project has been deposited at DDBJ/EMBL/GenBank under the accession number LFCU00000000.

Table 2 Genome sequencing project information

Growth conditions and genomic DNA preparation

The cell was retrieved directly from the field, added to the sample buffer of the illustra GenomiPhi V2 kit (GE Healthcare Life Sciences, Pittsburgh, PA), crushed manually with a sterile needle, heated for 3 min at 95 °C, and supplemented with the remaining ingredients for the MDA reaction [35]. Purity of the MDA product was assessed by amplifying the 16S rRNA gene sequence and directly sequencing the PCR product with Sanger. The genome was then reamplified with the illustra GenomiPhi HY DNA Amplification kit to yield enough material for whole genome sequencing.

Genome sequencing and assembly

The MDA product was sequenced with illumina MiSeq v2 technology at the Cornell University Institute of Biotechnology, Ithaca, NY. This resulted a total of 30,190,768 reads, which were quality checked, trimmed, and artifact/contamination filtered with DUK, a filtering program developed at the JGI that removes known Illumina sequencing and library preparation artifacts. Additionally, reads were screened for human, cat, and dog contaminant sequences. The remaining 29,696,136 reads were passed to SPAdes [36] and assembled into 586 contigs >2 kb, representing 7,614,708 bp. This dataset was uploaded in IMG/mer [37] under analysis project number Ga0064002, and further decontaminated manually. Only contigs affiliating with the Thiotrichaceae / Chromatiales lineage were finally uploaded in IMG/er [38] under analysis project number Ga0065144. This final dataset is the draft genome of Candidatus A. palustre and consists of 3,645,683 bp on 276 contigs, and the coverage is 375x. CheckM is software that is designed to assess quality and completeness of (meta)genomes [34], and our analysis of the draft genome dataset revealed a completeness of 83.36 % based on the finding of 503/538 lineage specific maker genes (marker lineage Gammaproteobacteria ), and a contamination value of 1.13 %, which is in the error range (≤6 %) of contamination estimates of incomplete (~70 %) genomes [34]. Strain heterogeneity, tested by the amino acid identity (AAI) between multi-copy genes [34], is 0.

Genome annotation

Gene calling and functional annotation was performed automatically by IMG [33, 39] during the upload process. We are currently manually verifying annotations of interest, constructing databases using Uniprot (Swissprot and TrEMBL) and blasting against these with the Achromatium draft genome using the integrated tblastn tool in IMG/er.

Genome properties

The Candidatus Achromatium palustre draft genome is 3,645,683 bp in size, and distributed on 276 contigs that are between 2012 and 57,118 bp in length. The N50 is 18,361 bp, and the G + C content is 38.08 %. Based on sequence comparison of nearly full-length 16S rRNA genes, the phylogenetic affiliation of the Candidatus Achromatium palustre genome is in cluster A among other Achromatium spp. sequences, including the three previously published draft genomes (Fig. 2). The Achromatium lineage is a sister lineage to the Chromatiaceae [2, 3, 8, 22, 24] containing purple sulfur bacteria such as Thiorhodovibrio and Chromatium (Fig. 2). IMG identified 3,400 genes, of which 3,343 encoded proteins (98.32 %), 57 encoded rRNA (1.68 %) and no pseudogenes (0.00 %). Among the 57 rRNA genes, one operon contained the 16S rRNA, 23S rRNA, and 5S rRNA gene. An additional truncated 5S rRNA gene was located on a different contig, and the sequence is identical to the full-length 5S rRNA gene. Furthermore, we find, e.g., 42 tRNA genes, genes for transcription and translation, DNA replication and repair, cell motility and chemotaxis. Details are given in Fig. 3, and Tables 3 and 4. We did not identify indications for plasmid DNA.

Fig. 3
figure 3

Graphical simulated circular genome of 276 concatenated contigs of the Candidatus A. palustre draft genome. The contigs were concatenated in Geneious 6.0.1 [50] using the random order of appearance in IMG, and the map was generated in Geneious and CGView [51]. The concatenated contigs are shown in blue, open reading frames (ORFs) in red in both directions, and the GC content in black

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

Further insights into the coding regions of the draft genome will be given elsewhere.

Conclusions

Details of Achromatium spp. genomes promise further insight into the ecophysiology of these unique organisms. The draft genome of Candidatus A plaustre is one of the first steps to unravel the phenotypic and physiological adaptations of Achromatium spp. occurring in different redox gradient systems as well as across divers salinities. A comparison with the brackish Achromatium genomes and prospect freshwater Achromatium spp. genomes, as well as with future metagenomes of different Achromatium -containing habitats, will be conducted and promise highly valuable information. Future analyses will not only include the investigation of nutrient pathways and modes of energy generation in these organisms, but also potential insights into calcium transport and calcite accumulation.

References

  1. Bersa E. Über das Vorkommen von kohlensaurem Kalk in einer Gruppe von Schwefelbakterien. Wien: Sitzungsbericht Akademie der Wissenschaften, mathematisch-naturwissenschaftliche Klasse, I Abteilung; 1920. p. 231–59.

    Google Scholar 

  2. Gray ND, Howarth R, Rowan A, Pickup RW, Jones JG, Head IM. Natural communities of Achromatium oxaliferum comprise genetically, morphologically, and ecologically distinct subpopulaitons. Appl Environ Microbiol. 1999;65(11):5089–99.

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Head IM, Gray ND, Clarke KJ, Pickup RW, Jones JG. The phylogenetic position and ultrastructure of the uncultured bacterium Achromatium oxaliferum. Microbiology. 1996;142:2341–54.

    Article  CAS  PubMed  Google Scholar 

  4. Nadson GA. Über Schwefelmikroorganismen des Hapsaler Meerbusens. Bulletin du Jardin Impériale Botanique de St-Pétersbourg. 1913;13:106–12.

    Google Scholar 

  5. Schewiakoff W. Über einen neuen bacterienähnlichen Organismus des Süsswassers. Heidelberg: University Heidelberg; 1892. p. 1–36.

    Google Scholar 

  6. Lackey JB, Lackey EW. The habitat and description of a new genus of sulphur bacterium. J Gen Microbiol. 1961;26:29–39.

    Article  CAS  PubMed  Google Scholar 

  7. Salman V, Yang T, Berben T, Klein F, Angert ER, Teske A. Calcite-accumulating large sulfur bacteria of the genus Achromatium in Sippewissett salt marsh. ISME J. 2015;9(11):2503–14.

    Article  CAS  PubMed  Google Scholar 

  8. Gray ND, Head IM. The family Achromatiaceae. In: Rosenberg E, Delong EF, Lory S, Stackebrandt E, Thompson FL, editors. The Prokaryotes: Gammaproteobacteria. Berlin-Heidelberg: Springer; 2014. p. 1–14.

    Google Scholar 

  9. West GS, Griffiths BM. The lime-sulphur bacteria of the genus Hillhousia. Ann Bot. 1913;27:83–91.

    Google Scholar 

  10. Gray ND, Comaskey D, Miskin IP, Pickup RW, Suzuki K, Head IM. Adaptation of sympatric Achromatium spp. to different redox conditions as a mechanism for coexistence of functionally similar sulphur bacteria. Environ Microbiol. 2004;6(7):669–77.

    Article  CAS  PubMed  Google Scholar 

  11. Gray ND, Howarth R, Pickup RW, Jones JG, Head IM. Substrate uptake by uncultured bacteria from the genus Achromatium determined by microautoradiography. Appl Environ Microbiol. 1999;65(11):5100–6.

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Gray ND, Howarth R, Pickup RW, Jones JG, Head IM. Use of combined microautoradioraphy and fluorescence in situ hybridization to determine carbon metabolism in mixed natural communities of uncultured bacteria from the genus Achromatium. Appl Environ Microbiol. 2000;66(10):4518–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Gray ND, Pickup RW, Jones JG, Head IM. Ecophysiological evidence that Achromatium oxaliferum is responsible for the oxidation of reduced sulfur species to sulfate in a freshwater sediment. Appl Environ Microbiol. 1997;63(5):1905–10.

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Head IM, Gray ND, Howarth R, Pickup RW, Clarke KJ, Jones JG. Achromatium oxaliferum - understanding the unmistakable. In: Schink B, editor. Advances in microbial ecology. Volume 16. New York: Kluwer Academic/Plenum Publishers; 2000. p. 1–40.

    Chapter  Google Scholar 

  15. Babenzien HD. Achromatium oxaliferum and its ecological niche. Zentralbl Mikrobiol. 1991;146:41–9.

    Google Scholar 

  16. Babenzien HD, Sass H. The sediment-water interface - habitat of the unusual bacterium Achromatium oxaliferum. Arch Hydrobiol Spec Issues Adv Limnol. 1996;48:247–51.

    Google Scholar 

  17. Mansor M, Hamilton T, Fantle MS, Macalady JL. Metabolic diversity and ecological niches of Achromatium populations revealed with single-cell genomic sequencing. Front Microbiol. 2015;6(822):1–14.

    Google Scholar 

  18. Schulz HN, Jørgensen BB. Big bacteria. Ann Rev Microbiol. 2001;55:105–37.

    Article  CAS  Google Scholar 

  19. de Boer WE, La Riviere JWM, Schmidt K. Some properties of Achromatium oxaliferum. Antonie Van Leeuwenhoek. 1971;37:553–63.

    Article  PubMed  Google Scholar 

  20. Seitz AP, Nielsen TH, Overmann J. Physiology of purple sulfur bacteria forming macroscopic aggregates in Great Sippewissett salt marsh, Massachusetts. FEMS Microbiol Ecol. 1993;12:225–36.

    Article  CAS  Google Scholar 

  21. Wilbanks EG, Jaekel U, Salman V, Humphrey PT, Eisen JA, Facciotti MT et al. A sulfurous symbiosis: microscale sulfur cycling in the pink berry consortia of the Sippewissett salt marsh. Environ Microbiol. 2014: doi:10.1111/1462-2920.12388.

  22. Howarth R, Unz RF, Seviour EM, Seviour RJ, Blackall LL, Pickup RW, et al. Phylogenetic relationships of filamentous sulfur bacteria (Thiothrix spp. and Eikelboom type 021N bacteria) isolated from wastewater-treatment plants and description of Thiothrix eikelboomii sp. nov., Thiothrix unzii sp. nov., Thiothrix fructosivorans sp. nov. and Thiothrix defluvii sp. nov. Int J Syst Bacteriol. 1999;49:1817–27.

    Article  CAS  PubMed  Google Scholar 

  23. Tindall BJ, Rossello-Mora R, Busse HJ, Ludwig W, Kampfer P. Notes on the characterization of prokaryote strains for taxonomic purposes. Int J Syst Evol Microbiol. 2010;60:249–66.

    Article  CAS  PubMed  Google Scholar 

  24. Glöckner FO, Babenzien HD, Wulf J, Amann R. Phylogeny and diversity of Achromatium oxaliferum. Syst Appl Microbiol. 1999;22(1):28–38.

    Article  PubMed  Google Scholar 

  25. Massart J. Recherches sur les organismes inferieur. Sur le protplame des Schizophytes. Section C. Schizomycetes, b. Thiobacterries Bruxelles: Univ de Bruxelles, tome V; 1901.

  26. Van Niel CB. Family A. Achromatiaceae Massart. In: Breed RS, Murray EGD, Hitchens AP, editors. Bergey’s Manual of Determinative Bacteriology. 6th ed. Baltimore: The Williams and Wilkins Company; 1948. p. 997–9.

    Google Scholar 

  27. Migula W. Ueber ein neues System der Bakterien. Arbeiten aus dem Bakteriologischen Institut der Technischen Hochschule zu Karlsruhe, Germany; 1894:235–38.

  28. Buchanan RE. Family III. Leucotrichaceae. In: Breed RS, Murray EGD, Smith NR, editors. Bergey’s Manual of Determinative Bacteriology. 7th ed. Baltimore: The Williams and Wilkins Company; 1957. p. 850–1.

    Google Scholar 

  29. Buchanan RE. Beggiatoales. In: Breed RS, Murray EGD, Smith NR, editors. Bergey’s Manual of Determinative Bacteriology. 7th ed. Baltimore: The Williams and Wilkins Company; 1957. p. 837–53.

    Google Scholar 

  30. Strohl WR. Order III. Beggiatoales. In: Staley JT, Bryant MP, Pfennig N, Holt JG, editors. Bergey’s Manual of Systematic Bacteriology. Volume 3. Baltimore: Williams & Wilkins Company; 1989. p. 2089–106.

    Google Scholar 

  31. Garrity GM, Bell JA, Lilburn T. Family I. Thiotrichaceae fam. nov. In: Garrity GM, Brenner DJ, Krieg NR, Staley JT, editors. Bergey’s Manual of Systematic Bacteriology, Volume 2. 2nd ed. New York: Springer; 2005. p. 131–78.

    Chapter  Google Scholar 

  32. Garrity GM, Bell JA, Lilburn T. Order V. Thiotrichales ord. nov. In: Garrity GM, Brenner DJ, Krieg NR, Staley JT, editors. Bergey’s Manual of Systematic Bacteriology, Volume 2. New York: Springer; 2005. p. 131–78.

    Chapter  Google Scholar 

  33. Markowitz VM, Chen I-MA, Palaniappan K, Chu K, Szeto E, Grechkin Y, et al. IMG: the integrated microbial genomes database and compartive analysis system. Nucleic Acids Res. 2012;40:D115–2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Parks DH, Imelfort M, Skennerton CT, Hugenholtz P, Tyson GW. Check M: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Peer J Pre Prints. 2014;No. e554v1:1–39.

    Google Scholar 

  35. Spits C, Le Caignec C, De Rycke M, Van Haute L, Van Steirteghem A, Liebaers I, et al. Whole-genome multiple displacement amplification from single cells. Nature Protocols. 2006;1(4):1965–70.

    Article  CAS  PubMed  Google Scholar 

  36. Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, et al. SPAdes: a new genome assembly algorithm and its aplications to single-cell sequencing. J Comput Biol. 2012;19(5):455–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Markowitz VM, Ivanova NN, Szeto E, Palaniappan K, Chu K, Dalevi D, et al. IMG/M: a data management and analysis system for metagenomes. Nucleic Acids Res. 2008;36:D534–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Markowitz VM, Mavromatis K, Ivanova NN, Chen I-MA, Chu K, Kyrpides N. IMG ER: a system for microbial genome annotation expert review and curation. Bioinformatics. 2009;25(17):2271–8.

    Article  CAS  PubMed  Google Scholar 

  39. Mavromatis K, Ivanova NN, Chen I-MA, Szeto E, Markowitz VM, Kyrpides N. The DOE-JGI standard operating procedure for the annotation of microbial genomes. Stand Genomic Sci. 2009;1:63–7.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Field D, Garrity G, Gray T, Morrison N, Selengut J, Sterk P, et al. The minimum information about a genome sequence (MIGS) specification. Nat Biotech. 2008;26(5):541–7.

    Article  CAS  Google Scholar 

  41. 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(12):4576–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Stackebrandt E, Murray EGD, Trüper HG. Proteobacteria classis nov., a name for the phylogenetic taxon that includes the purple bacteria and their relatives. Int J Syst Bacteriol. 1988;38(3):321–5.

    Article  Google Scholar 

  43. Garrity GM, Lilburn T. Phylum XIV. Proteobacteria phyl. nov. In: Garrity GM, Brenner DJ, Krieg NR, Staley JT, editors. Bergey’s Manual of Systematic Bacteriology. Volume 2. New York: Springer; 2005.

    Google Scholar 

  44. Euzeby J. Validation of publication of new names and new combinations previously effectively published outside the IJSEM. Int J Syst Evol Microbiol. 2005;55:2235–8.

    Article  Google Scholar 

  45. Garrity GM, Bell JA, Lilburn T. Class III. Gammaproteobacteria class. nov. In: Garrity GM, Brenner DJ, Krieg NR, Staley JT, editors. Bergey’s Manual of Systematic Bacteriology. Volume 2. New York: Springer; 2005.

    Google Scholar 

  46. Skerman VBD, McGowan V, Sneath PHA. Approved lists of bacterial names. Int J Syst Bacteriol. 1980;30:225–420.

    Article  Google Scholar 

  47. Murray RGE, Stackebrandt E. Taxonomic note - implementation of the provisional status Candidatus for incompletely described prokaryotes. Int J Syst Bacteriol. 1995;45(1):186–7.

    Article  CAS  PubMed  Google Scholar 

  48. Ashburner M, Ball CA, Blake JA, Botstein D, Butler H, Cherry JM, et al. Gene ontology: tool for the unification of biology. Nat Genetics. 2000;25:25–9.

    Article  CAS  PubMed  Google Scholar 

  49. Ludwig W, Strunk O, Westram R, Richter L, Meier H, Yadhukumar, et al. ARB: a software environment for sequence data. Nucleic Acids Res. 2004;32(4):1363–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M, Sturrock M, et al. Geneious Basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics. 2012;28(12):1647–9.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Grant JR, Stothard P. The CGView Server: a comparative genomics tool for circular genomes. Nucleic Acids Res. 2008;36:181–4.

    Article  Google Scholar 

Download references

Acknowledgements

Cell collection was financially assisted by the Marine Biological Laboratories and the Horace W Stunkard Scholarship Fund. Sequencing was funded by the NSF MCB 1244378. VS was supported by the Deutsche Forschungsgemeinschaft (Sa 2505/1-1) and NSF IOS 1354911. TB was supported by the ERC advanced GrantPARASOL (No. 322551). The work conducted by the U.S. Department of Energy Joint Genome Institute, a DOE Office of Science User Facility, is supported under Contract No. DE-AC02-05CH11231.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Verena Salman.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

VS designed the study, carried out the molecular genetic studies, performed bioinformatic analysis, and drafted the manuscript. RB carried out the qc and sequence alignment and performed further bioinformatic analysis. TB performed bioinformatic analysis, including ongoing verification of annotations. TW contributed valuable assistance regarding the performance of sequencing and sequence analysis. AT and EA conceived of the study, participated in its design and coordination and helped to draft the manuscript. All authors read and approved the final manuscript.

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

Salman, V., Berben, T., Bowers, R.M. et al. Insights into the single cell draft genome of “Candidatus Achromatium palustre”. Stand in Genomic Sci 11, 28 (2016). https://doi.org/10.1186/s40793-016-0146-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40793-016-0146-x

Keywords