Skip to main content

Complete genome sequence of Ferroglobus placidus AEDII12DO

Abstract

Ferroglobus placidus belongs to the order Archaeoglobales within the archaeal phylum Euryarchaeota. Strain AEDII12DO is the type strain of the species and was isolated from a shallow marine hydrothermal system at Vulcano, Italy. It is a hyperthermophilic, anaerobic chemolithoautotroph, but it can also use a variety of aromatic compounds as electron donors. Here we describe the features of this organism together with the complete genome sequence and annotation. The 2,196,266 bp genome with its 2,567 protein-coding and 55 RNA genes was sequenced as part of a DOE Joint Genome Institute Laboratory Sequencing Program (LSP) project.

Introduction

Strain AEDII12DO (=DSM 10642) is the type strain of the species Ferroglobus placidus. It was isolated from a shallow hydrothermal vent system at Vulcano Island, Italy [1]. F. placidus is metabolically quite versatile. It was isolated based on its ability to use ferrous iron as an electron donor, and was also shown to use hydrogen and sulfide as electron donors, with nitrate or thiosulfate as electron acceptors [1]. Subsequently, it was shown to produce N2O from nitrite, which is an unusual ability for an anaerobic organism [2]. It can also oxidize acetate and several aromatic compounds using ferric iron as the electron acceptor [3,4]. F. placidus is the first archaeon found to anaerobically oxidize aromatic compounds [4]. The genes and pathways involved in degradation of benzene, benzoate, and phenol have been recently characterized [5,6].

F. placidus is the only species in the genus Ferroglobus. It belongs to the family Archaeoglobaceae, which also contains the genera Archaeoglobus and Geoglobus. Genome sequences have been published for A. fulgidus and A. profundus [7,8]. Figure 1 shows the phylogenetic relationships between members of the family Archaeoglobaceae.

Figure 1.
figure 1

16S ribosomal RNA phylogenetic tree of Archaeoglobaceae. The tree was generated with weighbor [9] through the Ribosomal Database Project website [10] and displayed with njplot [11]. Organisms with two asterisks after the name are those with complete genomes sequenced and published. Those with one asterisk have a genome project in progress, according to the Genomes OnLine Database [12]. Methanocaldococcus jannaschii is the outgroup.

Organism information

F. placidus was isolated from a mixture of sand and water at a beach close to Vulcano Island, Italy [1]. The sample was taken from a depth of 1 m; the temperature of the sample was 95°C and the pH was 7.0 [1]. A 1.0 mL aliquot of the sample was incubated in FM medium at 85°C with shaking. The medium contained FeS as an electron donor [1]. F. placidus was isolated from the enrichment culture using optical tweezers [1]. The cells are irregular cocci with a triangular shape, and one or two flagella were present [1]. Growth occurred between 65°C and 95°C with an optimum of 85°C [1]. The optimal pH for growth was 7.0, and growth was observed over a range of 6.0 ­to 8.5 [1]. The optimal salinity for growth was 2.0%, with growth occurring between 0.5 and 4.5% NaCl [1]. F. placidus could use ferrous iron, hydrogen, or sulfide as electron donors and nitrate or thiosulfate as electron acceptors [1]. F. placidus also can anaerobically oxidize aromatic compounds with ferric iron as electron acceptor. The aromatic compounds it can utilize include benzene, benzoate, phenol, 4-hydroxybenzoate, benzaldehyde, p-hydroxybenzaldehyde and t-cinnamic acid [4,5]. The features of the organism are listed in Table 1.

Table 1. Classification and general features of F. placidus AEDII12DO according to the MIGS recommendations [13].

Genome sequencing information

Genome project history

This organism was selected for sequencing based on its phylogenetic position and its phenotypic differences from other members of the family Archaeoglobaceae. It is part of a Laboratory Sequencing Program (LSP) project to sequence diverse archaea. The genome project is listed in the Genomes On Line Database [12] and the complete genome sequence has been deposited in GenBank. Sequencing, finishing, and annotation were performed by the DOE Joint Genome Institute (JGI). A summary of the project information is shown in Table 2.

Table 2. Genome sequencing project information

Growth conditions and DNA isolation

The strain Ferroglobus placidus AEDII12DO (containing plasmid XY) has been deposited in the Deutsche Sammlung von Mikroorganismen und Zellkulturen (DSMZ) by Prof. Dr. K. O. Stetter, Lehrstuhl für Mikrobiologie, Universität Regensburg, Universitätsstr. 31, D-93053 Regensburg, Germany as DSM 10642.

F. placidus strain AEDII12DO was obtained from the DSMZ. Strict anaerobic culturing and sampling techniques were used throughout [22,23]. Ten 100 ml bottles of F. placidus cells were grown with acetate (10 mM) as the electron donor, and Fe(III) citrate (56 mM) as the electron acceptor. F. placidus medium was prepared as previously described [4]. After autoclaving, FeCl2 (1.3 mM), Na2SeO4 (30 µg/L), Na2WO4 (40 µg/L), APM salts (1 g/L MgCl2, 0.23 g/L CaCl2), DL vitamins [24] and all electron donors were added to the sterilized medium from anaerobic stock solutions. Cultures were incubated under N2:CO2 (80:20) at 85 °C in the dark.

For extraction of DNA, cultures (100 ml in 156 ml serum bottles) were divided into 50 ml conical tubes (Falcon), and cells were pelleted by centrifugation at 3,000 x g for 20 minutes. Cell pellets were resuspended in 10 ml TE sucrose buffer (10 mM Tris, pH 8.0, 1 mM EDTA, and 6.7% sucrose). The resuspended cells were distributed into 10 different 2 ml screw cap tubes and 3 µl Proteinase K (20 mg/ml), 30 µl sodium dodecyl sulfate (10% solution), and 10 µl RNase A (5 ug/ul) were added to each tube. Tubes were incubated at 37ºC for 20 min, and centrifuged at 16,100 x g for 15 minutes. The supernatant was transferred to a new set of tubes and 600 µl phenol (TE saturated, pH 7.3), and 400 µl chloroform-isoamyl alcohol were added. These tubes were then mixed on a Labquake rotator (Barnstead/Thermolyne, Dubuque, Iowa) for 10 min and centrifuged at 16,100 x g for 5 min. The aqueous layer was removed and transferred to new 2-ml screw cap tubes. The phenol/chloroform extraction step was performed again. The aqueous layer was transferred to a new tube, and 100 µl 5 M ammonium acetate, 20 µl glycogen (5 mg/ml; Ambion), and 1 ml cold (-20 °C) isopropanol (Sigma) were added. Nucleic acids were precipitated at −30 °C for 1 hour and pelleted by centrifugation at 16,100 x g for 30 min. The pellet was then cleaned with cold (-20 °C) 70% ethanol, dried, and resuspended in sterile nuclease-free water (Ambion).

Genome sequencing and assembly

The genome of F. placidus was sequenced at the Joint Genome Institute using a combination of Illumina and 454 technologies. An Illumina GAII shotgun library with reads of 539 Mb, a 454 Titanium draft library with average read length of 292.3 bases, and a paired-end 454 library with an average insert size of 15.5 Kb were generated for this genome. All general aspects of library construction and sequencing performed at the JGI can be found at the DOE JGI website [25].

Illumina sequencing data was assembled with Velvet [26], and the consensus sequences were shredded into 1.5 kb overlapped fake reads and assembled together with the 454 data. Draft assemblies were based on 104 Mb 454 draft data and 454 paired end data. The initial Newbler assembly contained 33 contigs in 1 scaffold. We converted the initial 454 assembly into a phrap assembly by making fake reads from the consensus, collecting the read pairs in the 454 paired end library.

The Phred/Phrap/Consed software package [27] was used for sequence assembly and quality assessment [28–30] in the following finishing process. After the shotgun stage, reads were assembled with parallel phrap (High Performance Software, LLC). Possible mis-assemblies were corrected with gapResolution (Cliff Han, unpublished), Dupfinisher [31], or sequencing cloned bridging PCR fragments with subcloning or transposon bombing (Epicentre Biotechnologies, Madison, WI). Gaps between contigs were closed by editing in Consed, by PCR and by Bubble PCR primer walks. A total of 43 additional reactions were necessary to close gaps and to raise the quality of the finished sequence.

Genome annotation

Genes were identified using Prodigal [32], followed by a round of manual curation using GenePRIMP [33]. The predicted CDSs were translated and used to search the National Center for Biotechnology Information (NCBI) nonredundant database, UniProt, TIGRFam, Pfam, PRIAM, KEGG, COG, and InterPro databases. The tRNAScanSE tool [34] was used to find tRNA genes, whereas ribosomal RNAs were found by using BLASTn against the ribosomal RNA databases. The RNA components of the protein secretion complex and the RNase P were identified by searching the genome for the corresponding Rfam profiles using INFERNAL [35]. Additional gene prediction analysis and manual functional annotation was performed within the Integrated Microbial Genomes (IMG) platform [36] developed by the Joint Genome Institute, Walnut Creek, CA, USA [37].

Genome properties

The genome includes one circular chromosome and no plasmids, for a total size of 2,196,266 bp (Table 3 and Figure 2). This genome size is almost the same as that of A. fulgidus and approximately 0.6 Mbp larger than that of A. profundus. The mol percent G+C is 44.1%, close to the values found in the Archaeoglobus genomes. A total of 2,622 genes were identified, 55 RNA genes and 2,567 protein-coding genes. There are 87 pseudogenes, comprising 3.4% of the protein-coding genes. The start codon is ATG in 83.7% of the genes, GTG in 12.2%, and TTG in 5.8%. This distribution is more similar to that of A. profundus, to which F. placidus is closely related (Figure 1), than to that of A. fulgidus. There is one copy of each ribosomal RNA. The 5S rRNA is not found adjacent to the 16S and 23S rRNAs. Table 4 shows the distribution of genes in COG categories.

Figure 2.
figure 2

Graphical circular map of the chromosome. From outside to the center: Genes on forward strand (colored by COG categories), genes on reverse strand (colored by COG categories), RNA genes (tRNAs green, rRNAs red, other RNAs black), GC content, and GC skew.

Table 3. Nucleotide content and gene count levels of the genome
Table 4. Number of genes associated with the 25 general COG functional categories

Insights from the genome

Nitrogen metabolism

Some aspects of the genome of F. placidus have been compared with those of A. fulgidus and A. profundus [8]. Here we will focus on some additional aspects of the F. placidus genome. F. placidus has been found to use nitrate as an electron acceptor and produces N2O with NO as an intermediate [2]. Genes likely to encode a nitrate reductase (Ferp_0311-0314) and an adjacent nitrate transporter (Ferp_0315) were identified. Based on the experimental results, F. placidus is expected to have a nitric oxide-forming nitrite reductase. There are two types of this protein: cytochrome cd1 type and copper type [38]. F. placidus appears to lack both of these types of nitrite reductase, so it may have a new version of this enzyme. F. placidus was found to produce N2O, and it has a NorBC-type nitric oxide reductase (Ferp_1340-1341). Surprisingly it also has a nitrous oxide reductase (Ferp_0128), suggesting that under some conditions F. placidus may carry out complete denitrification from nitrate to N2.

Central metabolism

F. placidus likely can not metabolize sugars as the Entner-Doudoroff pathway is absent from its genome, and the critical rate-limiting enzyme in the glycolysis pathway, 6-phosphofructokinase, also could not be identified. A complete gluconeogenesis pathway is present (Figure 3), including the recently discovered archaeal bifunctional fructose bisphosphate aldolase/phosphatase (Ferp_1532) [39]. A second fructose bisphosphate phosphatase may be present (Ferp_0896). Biosynthesis of C5 sugars for anabolic purposes proceeds through the reverse ribulose monophosphate pathway [40,41], in which fructose 6-phosphate is converted to hexulose 6-phosphate, from which formaldehyde is cleaved and ribulose 5-phosphate is generated.

Figure 3.
figure 3

Central metabolism of the hyperthermophilic archaeon Ferroglobus placidus. Abbreviations: G6PI: glucose-6-phosphate isomerase. FBPase: fructose-1,6-bisphosphatase. FBPA: fructose-1,6-bisphosphate aldolase. G3PDH: glycerol-3-phosphate dehydrogenase. TPI: triose-phosphate isomerase. PGK: phosphoglycerate kinase. PGM: phosphoglycerate mutase. PEP: phosphoenolpyruvate. PEPC: phosphoenolpyruvate carboxylase. PK: pyruvate kinase. PPDK: phosphoenolpyruvate dikinase. OadC: oxaloacetate decarboxylase. POR: pyruvate ferredoxin reductase (pyruvate synthase). ACL: acetyl-CoA ligase. AP: acetate phosphatase. CS: citrate synthase. ACN: aconitase. ICDH: isocitrate dehydrogenase. AKGD: alpha-ketoglutarate dehydrogenase. SuCoAS: succinyl-CoA synthase. SDH: succinate dehydrogenase. FUM: fumarase. MDH: malate dehydrogenase. PHI: 3-hexulose-6-phosphate isomerase. HPS: 3-hexulose-6-phosphate synthase. RPI: ribose-5-phosphate isomerase. FDH: formate dehydrogenase. MF: methanofuran. FMFDH: formylmethanofuran dehydrogenase. THM: tetrahydromethanopterin. THMFT: tetrahydromethanopterin formyl transferase. MTHMC: methenyltetrahydromethanopterin cyclohydrolase. MTHMR: methylentetrahydromethanopterin reductase. CODH/ACS: CO dehydrogenase/acetyl-CoA synthase. Gene designations in colors indicates association in clusters.

Similar to Archaeoglobus species, F. placidus is capable of autotrophic growth. The genome contains a gene coding for the large subunit of ribulose-1,5-bisphosphate carboxylase/oxygenase (RubisCO, Ferp_1506), which fixes CO2 in photosynthetic organisms, but many other enzymes of the Calvin-Benson cycle are missing. F. placidus probably uses RubisCO as part of an AMP recycling pathway [42] rather than for carbon fixation. F. placidus also contains the complete acetyl-CoA reductive pathway. Based on experimental results it was predicted to use this pathway for carbon fixation [2]. This pathway is composed of a methyl branch that reduces CO2 into a methyl group by a sequence of reactions similar to those found in methanogenesis (Fig. 3, inset), and a carbonyl branch that converts a second CO2 molecule into a carbonyl group. The two moieties are then joined to form acetyl-CoA.

Interestingly, there are two full copies of pyruvate ferredoxin oxidoreductase (POR, Ferp_0892-95 and Ferp_1823-26, 32–42% identical/47-60% similar), which generates pyruvate from acetyl-CoA and CO2. Conversely, the genome does not contain genes coding for the pyruvate dehydrogenase complex. All of the enzymes that comprise the TCA cycle could be accounted for, with the exception of a typical aconitase. However, two genes annotated as homoaconitate hydratase (Ferp_0702 and Ferp_2485) are 40% similar to the characterized aconitase from the thermoacidophilic archaeon Sulfolobus acidocaldarius [43]. Also F. placidus has the two subunits of a predicted aconitase (Ferp_0107-0108) [44].

Even though the genes involved in central metabolism are typically scattered in the genome, it is worth noting that many of these genes are grouped in clusters in F. placidus. For instance, the genes coding for the formylmethanofuran dehydrogenase (FMFDH, Ferp_0601-04) are located near the methenyltetrahydromethanopterin cyclohydrolase gene (MTHMC, Ferp_0606) from the reductive acetyl-CoA pathway. Similarly, two subunits of FMFDH (Ferp_0728-29), the methylenetetrahydromethanopterin reductase gene (MTHMR, Ferp_0743), the whole CO dehydrogenase/acetyl-CoA synthase operon (CODH/ACS, Ferp_0731-33, Ferp_0735-36), and the pyruvate kinase gene (PK, Ferp_0744), are in close proximity. The operon that contains the genes coding for one of the POR complexes (Ferp_0892-96) also includes genes coding for other enzymes that belong to central metabolism, such as one of the FBPases (Ferp_0896), an ATP-NAD kinase (Ferp_0897), and shikimate dehydrogenase (Ferp_0898), which participates in the biosynthesis of aromatic amino acids.

References

  1. Hafenbradl D, Keller M, Dirmeier R, Rachel R, Rossnagel P, Burggraf S, Huber H, Stetter KO. Ferroglobus placidus gen. nov., sp. nov., a novel hyperthermophilic archaeum that oxidizes Fe2+ at neutral pH under anoxic conditions. Arch Microbiol 1996; 166:308–314. PubMed doi:10.1007/s002030050388

    Article  CAS  PubMed  Google Scholar 

  2. Vorholt JA, Hafenbradl D, Stetter KO, Thauer RK. Pathways of autotrophic CO2 fixation and of dissimilatory nitrate reduction to N2O in Ferroglobus placidus. Arch Microbiol 1997; 167:19–23. PubMed doi:10.1007/s002030050411

    Article  CAS  PubMed  Google Scholar 

  3. Tor JM, Kashefi K, Lovley DR. Acetate oxidation coupled to Fe(III) reduction in hyperthermophilic microorganisms. Appl Environ Microbiol 2001; 67:1363–1365. PubMed doi:10.1128/AEM.67.3.1363-1365.2001

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  4. Tor JM, Lovley DR. Anaerobic degradation of aromatic compounds coupled to Fe(III) reduction by Ferroglobus placidus. Environ Microbiol 2001; 3:281–287. PubMed doi:10.1046/j.1462-2920.2001.00192.x

    Article  CAS  PubMed  Google Scholar 

  5. Holmes DE, Risso C, Smith JA, Lovley DR. Anaerobic oxidation of benzene by the hyperthermophilic archaeon Ferroglobus placidus. Appl Environ Microbiol 2011; 77:5926–5933. PubMed doi:10.1128/AEM.05452-11

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  6. Holmes DE, Risso C, Smith JA, Lovley DR. Genome-scale analysis of anaerobic benzoate and phenol metabolism in the hyperthermophilic archaeon Ferroglobus placidus. ISME J 2011; (In press). PubMed doi:10.1038/ismej.2011.88

  7. Klenk HP, Clayton RA, Tomb JF, White O, Nelson KE, Ketchum KA, Dodson RJ, Gwinn M, Hickey EK, Peterson JD, et al. The complete genome sequence of the hyperthermophilic, sulphate-reducing archaeon Archaeoglobus fulgidus. Nature 1997; 390:364–370. PubMed doi:10.1038/37052

    Article  CAS  PubMed  Google Scholar 

  8. von Jan M, Lapidus A, Glavina Del Rio T, Copeland A, Tice H, Cheng JF, Lucas S, Chen F, Nolan M, Goodwin L, et al. Complete genome sequence of Archaeoglobus profundus type strain (AV18). Stand Genomic Sci 2010; 2:327–346. PubMed doi:10.4056/sigs.942153

    Article  Google Scholar 

  9. Bruno WJ, Socci ND, Halpern AL. Weighted neighbor joining: a likelihood-based approach to distance-based phylogeny reconstruction. Mol Biol Evol 2000; 17:189–197. PubMed

    Article  CAS  PubMed  Google Scholar 

  10. Cole JR, Wang Q, Cardenas E, Fish J, Chai B, Farris RJ, Kulam-Syed-Mohideen AS, McGarrell DM, Marsh T, Garrity GM, et al. The Ribosomal Database Project: improved alignments and new tools for rRNA analysis. Nucleic Acids Res 2009; 37:D141–D145. PubMed doi:10.1093/nar/gkn879

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  11. Perrière G, Gouy M. WWW-query: an on-line retrieval system for biological sequence banks. Biochimie 1996; 78:364–369. PubMed doi:10.1016/0300-9084(96)84768-7

    Article  PubMed  Google Scholar 

  12. Liolios K, Chen IM, Mavromatis K, Tavernarakis N, Hugenholtz P, Markowitz VM, Kyrpides NC. The Genomes On Line Database (GOLD) in 2009: status of genomic and metagenomic projects and their associated metadata. Nucleic Acids Res 2010; 38:D346–D354. PubMed doi:10.1093/nar/gkp848

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  13. Field D, Garrity G, Gray T, Morrison N, Selengut J, Sterk P, Tatusova T, Thomson N, Allen MJ, Angiuoli SV, et al. The minimum information about a genome sequence (MIGS) specification. Nat Biotechnol 2008; 26:541–547. PubMed doi:10.1038/nbt1360

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  14. Woese CR, Kandler O, Wheelis ML. Towards a natural system of organisms: proposal for the domains Archaea, Bacteria, and Eucarya. Proc Natl Acad Sci USA 1990; 87:4576–4579. PubMed doi:10.1073/pnas.87.12.4576

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  15. Garrity GM, Holt JG. Phylum AII. Euryarchaeota phy. nov. In: Bergey’s Manual of Systematic Bacteriology, vol 1. 2nd ed. Edited by: Garrity GM, Boone DR, Castenholz RW. Springer, New York; 2001; pp 211–355.

    Chapter  Google Scholar 

  16. Garrity GM, Holt JG. Class VI. Archaeoglobi class. nov. In: Bergey’s Manual of Systematic Bacteriology, vol 1. The Archaea and the deeply branching and phototrophic Bacteria. Springer, New York; 2001; pp 349.

    Google Scholar 

  17. List Editor. Validation List no. 85. Validation of publication of new names and new combinations previously effectively published outside the IJSEM. Int J Syst Evol Microbiol 2002; 52:685–690. PubMed doi:10.1099/ijs.0.02358-0

  18. Huber H, Stetter KO. Order I. Archaeoglobales ord. nov. In: Bergey’s Manual of Systematic Bacteriology, vol. 1. The Archaea and the deeply branching and phototrophic Bacteria. Springer, New York; 2001; pp 349.

    Google Scholar 

  19. Huber H, Stetter KO. Family I. Archaeoglobaceae fam. nov. In: Bergey’s Manual of Systematic Bacteriology, vol. 1. The Archaea and the deeply branching and phototrophic Bacteria. Springer, New York; 2001; pp 349.

    Google Scholar 

  20. List Editor. Validation List no. 61. Validation of the publication of new names and new combinations previously effectively published outside the IJSB. Int J Syst Bacteriol 1997; 47:601–602. doi:10.1099/00207713-47-2-601

  21. Ashburner M, Ball CA, Blake JA, Botstein D, Butler H, Cherry JM, Davis AP, Dolinski K, Dwight SS, Eppig JT, et al. Gene ontology: tool for the unification of biology. The Gene Ontology Consortium. Nat Genet 2000; 25:25–29. PubMed doi:10.1038/75556

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  22. Balch WE, Fox GE, Magrum LJ, Woese CR, Wolfe RS. Methanogens: reevaluation of a unique biological group. Microbiol Rev 1979; 43:260–296. PubMed

    PubMed Central  CAS  PubMed  Google Scholar 

  23. Miller TL, Wolin MJ. A serum bottle modification of the Hungate technique for cultivating obligate anaerobes. Appl Microbiol 1974; 27:985–987. PubMed

    PubMed Central  CAS  PubMed  Google Scholar 

  24. Lovley DR, Phillips EJ. Novel mode of microbial energy metabolism: organic carbon oxidation coupled to dissimilatory reduction of iron or manganese. Appl Environ Microbiol 1988; 54:1472–1480. PubMed

    PubMed Central  CAS  PubMed  Google Scholar 

  25. Sequencing protocols. http://www.jgi.doe.gov/sequencing/protocols/prots_production.html

  26. Zerbino DR, Birney E. Velvet: algorithms for de novo short read assembly using de Bruijn graphs. Genome Res 2008; 18:821–829. PubMed doi:10.1101/gr.074492.107

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  27. The Phred/Phrap/Consed software package. http://www.phrap.com

  28. Ewing B, Hillier L, Wendl MC, Green P. Base-calling of automated sequencer traces using phred. I. Accuracy assessment. Genome Res 1998; 8:175–185. PubMed

    Article  CAS  PubMed  Google Scholar 

  29. Ewing B, Green P. Base-calling of automated sequencer traces using phred. II. Error probabilities. Genome Res 1998; 8:186–194. PubMed

    Article  CAS  PubMed  Google Scholar 

  30. Gordon D, Abajian C, Green P. Consed: a graphical tool for sequence finishing. Genome Res 1998; 8:195–202. PubMed

    Article  CAS  PubMed  Google Scholar 

  31. Han C, Chain P. Finishing repeat regions automatically with Dupfinisher. In Proceedings of the 2006 international conference on bioinformatics and computational biology, ed. Arabnia HR, Valafar H. CSREA Press, 2006:141–146.

  32. Hyatt D, Chen GL, Lacascio PF, Land ML, Larimer FW, Hauser LJ. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinformatics 2010; 11:119. PubMed doi:10.1186/1471-2105-11-119

    Article  PubMed Central  PubMed  Google Scholar 

  33. Pati A, Ivanova NN, Mikhailova N, Ovchinnikova G, Hooper SD, Lykidis A, Kyrpides NC. Gene PRIMP: a gene prediction improvement pipeline for prokaryotic genomes. Nat Methods 2010; 7:455–457. PubMed doi:10.1038/nmeth.1457

    Article  CAS  PubMed  Google Scholar 

  34. Lowe TM, Eddy SR. tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res 1997; 25:955–964. PubMed doi:10.1093/nar/25.5.955

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  35. INFERNAL software package. http://infernal.janelia.org

  36. DOE Joint Genome Institute. http://img.jgi.doe.gov

  37. Markowitz VM, Mavromatis K, Ivanova NN, Chen IMA, Chu K, Kyrpides NC. IMG ER: a system for microbial genome annotation expert review and curation. Bioinformatics 2009; 25:2271–2278. PubMed doi:10.1093/bioinformatics/btp393

    Article  CAS  PubMed  Google Scholar 

  38. Cutruzzolà F, Rinaldo S, Castiglione N, Giardina G, Pecht I, Brunori M. Nitrite reduction: a ubiquitous function from a pre-aerobic past. Bioessays 2009; 31:885–891. PubMed doi:10.1002/bies.200800235

    Article  PubMed  Google Scholar 

  39. Say RF, Fuchs G. Fructose 1,6-bisphosphate aldolase/phosphatase may be an ancestral gluconeogenic enzyme. Nature 2010; 464:1077–1081. PubMed doi:10.1038/nature08884

    Article  CAS  PubMed  Google Scholar 

  40. Soderberg T. Biosynthesis of ribose-5-phosphate and erythrose-4-phosphate in archaea: a phylogenetic analysis of archaeal genomes. Archaea 2005; 1:347–352. PubMed doi:10.1155/2005/314760

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  41. Orita I, Sato T, Yurimoto H, Kato N, Atomi H, Imanaka T, Sakai Y. The ribulose monophosphate pathway substitutes for the missing pentose phosphate pathway in the archaeon Thermococcus kodakaraensis. J Bacteriol 2006; 188:4698–4704. PubMed doi:10.1128/JB.00492-06

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  42. Sato T, Atomi H, Imanaka T. Archaeal type III RuBisCOs function in a pathway for AMP metabolism. Science 2007; 315:1003–1006. PubMed doi:10.1126/science.1135999

    Article  CAS  PubMed  Google Scholar 

  43. Uhrigshardt H, Walden M, John H, Anemüller S. Purification and characterization of the first archaeal aconitase from the thermoacidophilic Sulfolobus acidocaldarius. Eur J Biochem 2001; 268:1760–1771. PubMed doi:10.1046/j.1432-1327.2001.02049.x

    Article  CAS  PubMed  Google Scholar 

  44. Makarova KS, Koonin EV. Filling a gap in the central metabolism of archaea: prediction of a novel aconitase by comparative-genomic analysis. FEMS Microbiol Lett 2003; 227:17–23. PubMed doi:10.1016/S0378-1097(03)00596-2

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The work conducted by the US Department of Energy Joint Genome Institute is supported by the Office of Science of the U.S. Department of Energy under Contract No. DE-AC02-05CH11231.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Iain Anderson.

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly credited. 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

Cite this article

Anderson, I., Risso, C., Holmes, D. et al. Complete genome sequence of Ferroglobus placidus AEDII12DO. Stand in Genomic Sci 5, 50–60 (2011). https://doi.org/10.4056/sigs.2225018

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.4056/sigs.2225018

Keywords