Skip to main content

Complete genome sequence of Thioalkalivibrio sp. K90mix

Abstract

Thioalkalivibrio sp. K90mix is an obligately chemolithoautotrophic, natronophilic sulfur-oxidizing bacterium (SOxB) belonging to the family Ectothiorhodospiraceae within the Gammaproteobacteria. The strain was isolated from a mixture of sediment samples obtained from different soda lakes located in the Kulunda Steppe (Altai, Russia) based on its extreme potassium carbonate tolerance as an enrichment method. Here we report the complete genome sequence of strain K90mix and its annotation. The genome was sequenced within the Joint Genome Institute Community Sequencing Program, because of its relevance to the sustainable removal of sulfide from wastewater and gas streams.

Introduction

Thioalkalivibrio sp. K90mix is an obligately chemolithoautotrophic SOxB using CO2 as a carbon source and reduced inorganic sulfur compounds as an energy source. It belongs to the genus Thioalkalivibrio. This genus represents a dominant SOxB type in soda lakes — extremely alkaline and saline habitats - and is the first example of an obligate chemolithoautotroph capable of growing in saturated sodium carbonate brines. It forms a monophyletic group within the family Ectothiorhodospiraceae of the Gammaproteobacteria. The genus currently includes nine validly published species [1] and around 70, yet uncharacterized strains that are extremely salt-tolerant and genetically different from the characterized isolates recovered from hypersaline soda lakes [24]. The members are slow growing obligate autotrophs, well adapted to hypersaline (up to salt saturation) and alkaline (up to pH 10.5) conditions. Members of the genus Thioalkalivibrio have versatile metabolic capabilities, including oxidation of reduced sulfur compounds [1,2], denitrification [5,6] and thiocyanate utilization [7,8].

Apart from playing an important role in the sulfur cycle of soda lakes, Thioalkalivibrio species also are being used for the sustainable removal of sulfide from wastewater and gas streams [9,10]. In this process hydrogen sulfide is absorbed to a high salt alkaline solution, which is subsequently transferred to a bioreactor in which Thioalkalivibrio spp. oxidize HS- to elemental sulfur. The produced biosulfur can then be used as a fertilizer or fungicide [9].

To get a comprehensive understanding of the molecular mechanism by which Thioalkalivibrio sp. K90mix oxidize sulfur compounds and adapts to extreme alkaline (up to pH 10.5) and hypersaline conditions (up to 4 M of Na+ or 3.6 M of K+) it is necessary to identify the genes that are involved in these adaptations. The most important issues in this are the mechanism of sulfide oxidation, carbon assimilation at high pH, and bioenergetic adaptation to high salt and high pH. Here we present a summary classification and a set of features for Thioalkalivibrio sp. K90mix together with the description of the genomic sequencing and annotation.

Classification and features

Because of limited solubility of sodium carbonates in the biodesulfurization process, we made a series of enrichment cultures with an increasing ratio of potassium to sodium carbonate (potassium carbonates have a 2–5 times higher solubility than sodium carbonates). Thioalkalivibrio sp. K90mix was isolated from a culture that was inoculated with a mixture of sediment samples from different hypersaline soda lakes and was grown at the maximal possible substitution of sodium for potassium, 3.6 M K+/0.4 M Na+ (90% substitution).

Thioalkalivibrio sp. K90mix has rod-shaped cells with a polar flagellum (Figure 1), that elongate at high concentrations of K+ (Figure 1b). The strain is obligately alkaliphilic with a pH optimum of 10 (Table 1). It can tolerate a salinity of 4.0 M total Na+, but has an optimum of 0.3 M, sulfide concentrations up to 1 mM and a temperature up to 40°C. It has a preference for carbonate and sulfate as counter-anions over chloride and, therefore must be called “natronophilic”, instead of “haloalkaliphilic”. It utilizes ammonia, nitrate and nitrite as a nitrogen source. On the basis of 16S rRNA gene sequencing the strain belongs to the genus Thioalkalivibrio within the Gammaproteobacteria with T. thiocyanoxidans and T. nitratis as the closest described species (Figure 2). Most of the yet undescribed Thioalkalivibrio isolates from hypersaline lakes of Siberia and Mongolia also belong to this core genetic cluster of the genus Thioalkalivibrio.

Figure 1.
figure 1

Phase contrast micrographs of the cell morphology of Thioalkalivibrio sp. K90mix grown at pH 10 and 4 M Na+ (a), or 3.6 M K+/0.4 M Na+ (b).

Figure 2.
figure 2

Phylogenetic tree based on 16S rRNA sequences showing the phylogenetic position of Thioalkalivibrio sp. K90mix. The sequence was aligned to sequences stored in the SILVA database using the SINA Webaligner [24]. Subsequently, the aligned sequences were imported into ARB [25], and a neighbor joining tree was constructed. Sequences of members from the Alphaproteobacteria were used as outgroup, but were pruned from the tree. The scale bar indicates 1% sequence difference.

Table 1. Classification and general features of Thioalkalivibrio sp. K90mix according to the MIGS recommendations [11].

Genome sequencing information

Genome project history

Strain K90mix was selected for sequencing in the 2007 Joint Genome Institute Community Sequencing Program, because of its relevance to bioremediation. A summary of the project information is presented in Table 2. The complete genome sequence was finished in February 2010. The GenBank accession numbers are NC_013889 and NC_013930 for the chromosome and plasmid, respectively. The genome project is listed in the Genome OnLine Database (GOLD) [26] as project Gc01217. Sequencing was carried out at the Joint Genome Institute (JGI) Finishing was done by JGI-Los Alamos National Laboratory (LANL) and initial automatic annotation by JGI-Oak Ridge National Laboratory (ORNL).

Table 2. Genome sequencing project information

Growth conditions and DNA isolation

Thioalkalivibrio sp. K90mix was grown with 40 mM thiosulfate as an energy source in standard sodium carbonate-bicarbonate medium at pH 10 and 2 M Na+ [2] at 35oC with shaking at 200 rpm. The cells were harvested by centrifugation and stored at minus 80°C for DNA extraction. Genomic DNA was obtained using phenol-chloroform-isoamylalcohol (PCI) extraction. The genomic DNA was extracted using PCI and precipitated with ethanol. The pellet was dried under vacuum and subsequently dissolved in water. The quality and quantity of the extracted DNA was evaluated using the DNA Mass Standard Kit provided by the JGI.

Genome sequencing and assembly

The genome of Thioalkalivibrio sp. K90mix was sequenced using a combination of Sanger and 454 sequencing platforms. All general aspects of library construction and sequencing can be found at the JGI website [27]. Pyrosequencing reads were assembled using the Newbler assembler version 1.1.02.15 (Roche). Large Newbler contigs were broken into 3,292 overlapping fragments of 1,000 bp and entered into assembly as pseudo-reads. The sequences were assigned quality scores based on Newbler consensus q-scores with modifications to account for overlap redundancy and adjust inflated q-scores. A hybrid 454/Sanger assembly was made using the PGA assembler. Possible mis-assemblies were corrected and gaps between contigs were closed by editing in Consed, by custom primer walks from sub-clones or PCR products. A total of 181 Sanger finishing reads were produced to close gaps, to resolve repetitive regions, and to raise the quality of the finished sequence. Illumina reads were used to improve the final consensus quality using an in-house developed tool (the ‘Polisher’ [28]). The error rate of the completed genome sequence is less than 1 in 100,000. Together, the combination of the Sanger and 454 sequencing platforms provided 42.1× coverage of the genome. The final assembly contains 28,443 Sanger reads (10.0×) and 419,015 pyrosequencing reads (32.1×).

Genome annotation

Genes were identified using Prodigal [29] as part of the Oak Ridge National Laboratory genome annotation pipeline followed by a round of manual curation using the JGI GenePRIMP pipeline [30]. 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. Additional gene prediction analysis and functional annotation was performed within the Integrated Microbial Genomes Expert Review (IMG-ER) platform [31].

Genome properties

The genome of strain K90mix consists of a circular chromosome with a size of 2.74 Mbp (Figure 3) and a linear plasmid of 240 Kbp. The G+C percentage determined from the genome sequence is 65.54%, which is similar to the value determined by thermal denaturation (65.8±0.5 mol%). There are 2942 genes of which 2888 are protein-coding genes and the remaining 54 are RNA genes. 33 pseudogenes were identified, constituting 1.12% of the total number of genes. The genome is smaller than that of “Thioalkalivibrio sulfidophilus” HL-EbGr7 [32], 2.98 Mbp versus 3.46 Mbp, but has a similar percentage of protein-coding genes (98.16% versus 98.06%). The properties and statistics of the genome are summarized in Table 3, and genes belonging to COG functional categories are listed in Table 4.

Figure 3.
figure 3

Graphical circular map of the chromosome of Thioalkalivibrio sp. strain K90mix. From outside to the center: Genes on the forward strand (Colored by COG categories), Genes on the reverse strand (colored by COG categories), RNA genes (tRNAs green, rRNAs red, other RNAs black), GC content, GC skew.

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

Insights from the genome sequence

Autotrophic growth

As mentioned before, [32] autotrophic growth at extremely high pH is a problem, because inorganic carbon is mainly present as carbonate (with bicarbonate as a minor fraction) at pH values above 10. This would demand active transport of bicarbonate into the cell. We found a gene related to stbA encoding a Na+/HCO3 symporter in the marine cyanobacterium Synechocystis sp. strain PCC 6803 [33]. Figure 4 shows a phylogenetic tree of different sequences related to StbA and the hydrophobicity profiles of StbA of Synechocystis sp. PCC 6803 and Thioalkalivibrio sp. K90mix. In addition, we have found genes for the large (TK90_0858) and small subunit (TK90_0859) of RuBisCO form 1Ac, and for the synthesis of α-carboxysomes (TK90_0860–TK90_0866), including csoSCA encoding a carboxysome shell alpha-type carbonic anhydrase, which was also found in genomic analysis of “Thioalkalivibrio sulfidophilus” HL-EbGr7 [32].

Figure 4.
figure 4

Phylogeny and hydropathy of StbA from different microorganisms. The blue boxes indicate sequences belonging to the cyanobacteria. The dots on the branches are bootstrap values between 50% and 75% (open dots) or between 75% and 100% (closed dots) The scale bar indicated 10% sequence difference. The sequence of Thioalkalivibrio sp. K90mix is indicated in bold. Hydropathy profiles were made for the StbA protein sequences of Synechocystis sp. PCC6803 and Thioalkalivibrio sp. K90mix to indicate transmembrane spanning domains.

Sulfur metabolism

Thioalkalivibrio sp. K90mix can oxidize sulfide/polysulfide, thiosulfate, sulfite (in vitro) and elemental sulfur to sulfate. Elemental sulfur is formed as an intermediate during sulfide and thiosulfate oxidation at oxygen limitation and near-neutral pH. Figure 5 shows a schematic overview of the different genes that are involved in the oxidation of sulfur compounds. The genome of Thioalkalivibrio sp. K90mix contains genes for flavocytochrome c/sulfide dehydrogenase (TK90_0236), which oxidizes sulfide to elemental sulfur. It contains an incomplete set of sox genes including soxYZ (TK90_0123 and TK90_0124), soxAX (TK90_0432 and TK90_0433) and two copies of soxB (TK90_0627 and TK90_1150), but is lacking soxCD, which would allow oxidizing the sulfane atom of thiosulfate to the state of elemental sulfur, but no further. However, it does not contain the reverse dissimilatory sulfite reduction pathway to oxidize sulfur to sulfite, which has been found in the genome of “Thioalkalivibrio sulfidophilus” HLEbGr7 [32]. Absence of dsr genes has also been found for the green sulfur bacterium Chloroherpeton thalassium that can oxidize sulfide to elemental sulfur, but subsequently can only oxidize the produced sulfur very slowly [34], probably due to the absence of dsr. Frigaard and Dahl [35] suggested that the presence of a RuBisCo-like protein (RLP) might be involved in sulfur oxidation [36]. Genes encoding for the RuBisCo-like protein were not found, nor were genes encoding sulfur dioxygenase or sulfur oxygenase-reductase, which can oxidize or disproportionate sulfur in several acidophilic bacteria and archaea [37]. However, we found a gene cluster encoding two sulfur transferases (rhd, TK90_0630; sirA, TK90_0631) and a heterodisulfide reductase complex (TK90_0632–TK90_0637) consisting of hdrA, hdrB, and hdrC (Figure 6). dsrE was missing in this cascade, but was present at 3 other places in the genome (TK_0511, TK_0639, TK90_1244).

Figure 5.
figure 5

Schematic overview of the different genes that are involved in the oxidation of sulfur compounds, although the role of the Hdr complex has not been proven yet. The genes encoding the reverse dissimilatory sulfite reductase (dsr), which are present in the genome of Thioalkalivibrio sulfidophilus, are absent in the genome of Thioalkalivibrio sp. K90mix.

Figure 6.
figure 6

Comparison of the hdr cluster of A. ferroooxidans ATCC 23270 (AF), Thioalkalivibrio sp. K90mix (K90) and Thioalkalivibrio sulfidophilus HL-EbGR7 (HL). The heterodisulfide reductase complex consists of the genes encoding HdrC1, HdrB1, HdrA, orf2, HdrC2 and HdrB2. The accessory proteins are Rhd, and TusA. DsrE was not found in this order, but was present at other places in the genome. The percentage of amino-acid similarity is indicated.

The Hdr complex plays a function in the energy metabolism of methanogens [38] and sulfate-reducing prokaryotes [39]. In methanogens, the enzyme complex catalyzes the reversible reduction of the disulfide (CoM-S-S-CoB) of the two methanogenic thiol-coenzymes, coenzyme M (CoM-SH) and coenzyme B (CoB-SH); in sulfate reducing microorganisms the substrate (X-S-S-X) is not known. Recently, the genes encoding the Hdr complex have also been detected in the genomes of the acidophilic sulfur oxidizing bacteria Acidithiobacillus ferrooxidans [40] and Acidithiobacillus caldus [41]. Quantrini and co-workers [40] hypothesized that Hdr, like the dissimilatory sulfite reductase (dsr), is working in reverse, whereby sulfur (i.e., sulfane atom) is oxidized to sulfite, and electrons are donated to the quinone pool. Although the role of the Hdr complex in the sulfur metabolism still has to be confirmed by expression analysis, Hdr genes have not yet been detected in the genomes of the neutrophilic chemolithotrophic sulfur-oxidizing bacteria Thiomicrospira crunogena and Thiobacillus denitrificans or of phototrophic sulfur-oxidizing bacteria Allochromatium vinosum and Halorhodospira halophila. The absence of a reverse dsr pathway in Thioalkalivibrio sp. K90mix might be the reason why it can only tolerate sulfide concentrations up to 1 mM. Sulfite can be oxidized further to sulfate, either directly by sulfite dehydrogenase (sorA; TK90_0686) or indirectly via adenosine-5’-phosphosulfate (APS) by APS reductase encoded by aprBA (TK90_0064 and TK90_0065) and ATP sulfurylase encoded by sat (TK90_0062) [42]. All these genes are present in the investigated genome.

Energy metabolism and pH homeostasis

At this time, it is not clear how Thioalkalivibrio sp. K90mix can withstand the harsh conditions of high pH and salinity. The difference between the pH of the environment (pH 10) and the pH in the cell (pH 8) causes a reversed ΔpH and consequently lowers the proton motive force (PMF). Therefore, Thioalkalivibrio requires a special molecular mechanisms to obtain enough energy for growth. It certainly needs this energy, because the production of osmolytes, to withstand the high concentrations of salts, costs 55 molecules of ATP for one molecule of glycine betaine, and 110 molecules of ATP for 1 molecule of sucrose [43]. In addition, the chemolithoautotrophic life style of CO2 fixation is energetically very expensive. The redox potential of the substrate couple S°/HS- (−260 mV) is more positive that the potential of NAD+/NADH (-340 mV) and therefore the direct reduction of NAD+ in order to supply reducing equivalents for CO2 fixation is not possible. Reverse electron transport is necessary in order to produce enough NADH, necessary for CO2 fixation, which costs extra energy. In addition, because of the large pH gradient over the cell membrane Thioalkalivibrio needs special mechanisms to keep the intracellular pH around neutral (pH homeostasis), which again is an energy requiring process.

The genome has revealed genes encoding similar proteins as those found for “Thioalkalivibrio sulfidophilus” HL-EbGr7 [32]. We found genes for a proton-driven F0F1-type ATP synthase (i.e., subunit A TK90_2593, B TK90_2591, and C TK90_2592of the F0 subcomplex, and subunit alpha (TK90_2589), beta (TK90_2587), gamma (TK90_2588), delta (TK90_2590), and epsilon (TK90_2586) of the F1 subcomplex), genes encoding the proton-translocating NADH dehydrogenase (nuoABCDEFGHIJKLMN) (TK90_0708 to TK90_0721), as well as the genes for a putative primary sodium pump Rnf [44] (rnfABCDGE) (TK90_1790 to TK90_1795). In addition, we found several genes encoding different secondary sodium-dependent pumps, such as the Na+/H+ antiporters NhaP (TK90_1831) and Mrp (mnhA-G) (TK90_0748 to TK90_0752), which according to Padan et al. [45] both play an essential role in alkaline pH homeostasis. In addition, we found genes encoding transporters belonging to the SulP family (TK90_0019, TK90_0897, TK90_0985). Transporters of this group could be involved in the low affinity, but high flux of bicarbonate uptake [46]. In addition, genes encoding the sodium-depending flagellar motor PomA/B (TK90_1180 and TK90_1181) are also present in the genome (see below for more details). As Thioalkalivibrio sp. K90mix can stand high concentrations of potassium, we also searched for K+-transporters and found genes encoding the following transporters: TrkA-C (TK90_0502), TrkA-N (TK90_2266) and TrkH (TK90_2267) that are part of the potassium uptake system [47].

Chemotaxis and motility

We found different genes encoding methyl-accepting chemotaxis sensory transducers (TK90_0580, TK90_0949, TK90_1402, TK90_2562, TK90_2397) that are involved in chemotaxis. One of these genes, Aer (TK90_0580), encodes a redox sensor involved in aerotaxis. In E. coli, Aer regulates the motility behavior in gradients of oxygen, redox potential and certain nutrients by interacting with the CheA-CheW complex. We found genes encoding several different proteins of this complex, CheA (TK90_1178), CheW (TK90_1183 and TK90_1184), CheY (TK90_1176), CheZ (TK90_1177), CheB (TK90_1179), CheV (TK90_0924) and CheR (TK90_0925). Chemotaxis consists of a complex cascade of different reactions: the redox sensor Aer senses a difference in redox potential induced by a change in the environmental oxygen concentration, which leads to the autophosphorylation of the histidine protein kinase CheA. CheA phosphorylates CheY, which will switch on the flagellar motor (see [48] for a detailed overview). CheW acts as an adaptor protein, while CheB, CheR, CheZ, and CheV are involved in feedback regulation.

Thioalkalivibrio sp. K90mix has all the genes that are indispensable for the production of flagellar proteins (FlgA, TK90_0923; FlgB, TK90_0926; FlgC, TK90_0927; FlgD, TK90_0928; FlgE, TK90_0929; FlgF, TK90_0930; FlgG, TK90_0931; FlgH, TK90_0932; FlgI, TK90_0933; FlgK, TK90_0935; FlgL, TK90_0936; FlgM, TK90_0922; FlgN, TK90_0921; FliC, TK90_1400, 1448, 1450; FliD, TK90_1447; FliE, TK90_1157; FliF, TK90_1158; FliG, TK90_1159; FliH, TK90_1160; FliI, TK90_1161; FliJ, TK90_1162; FliK, TK90_1163; FliM, TK90_1165; FliN, TK90_1166; FliO, TK90_1167; FliP, TK90_1168; FliQ, TK90_1169; FliR, TK90_1170; FliS, TK90_1446; FlhA, TK90_1172; FlhB, TK90_1171; see Table 1 in [49]). Phylogenetic analysis of sequences encoding different flagellar motors showed, with significant bootstrap values, genes encoding both proton-driven motors (TK90_0578, TK90_0577) related to E. coli MotA and MotB, as well as sodium-driven motors (TK90_1180, TK1181) related to Vibrio cholera PomA and PomB and Bacillus subtilis MotP and MotS (Figure 7). The results also show the presence of proton- and sodium-driven flagellar motors in the Ectothiorhodospiraceae, “Thioalkalivibrio sulfidophilus”, Halorhodospira halophila, and Alkalilimnicola ehrlichei, as well as in Achromatium vinosum, Acidithiobacillus caldus, Vibrio alginolyticus and Vibrio parahaemolyticus. Vibrio cholerae, Bacillus subtilis, B. halodurans, and Nitrococcus mobilis only have sodium-driven flagellar motors. Grouping of the proton-driven MotA and sodium-driven PomA has also been found by Krulwich et al. [50].

Figure 7.
figure 7

Phylogenetic tree based on protein sequences of different flagellar motors. The sequences of the proton-driven flagellar motor MotAB from E. coli (ECDH10B_2031, ECDH10B_2030) and of the sodium-driven flagellar motor PomAB from V. cholerae (VC0892, VC0893) were used as reference proteins. Other proteins were selected after BLAST analysis. Subsequently, the selected protein sequences were aligned used Clustal W, and a neighbor joining tree was drawn using MEGA 5. Ac, Acidithiobacillus caldus ATCC 51756; Ae, Alkalilimnicola ehrlichei MLHE-1; Av, Allochromatium vinosum DSM 180; Bh, Bacillus halodurans C125; Bs, Bacillus subtilis strain 168; Ec, Escherichia coli DH10B; Hh, Halorhodospira halophila SL1; Nm, Nitrococcus mobilis NB-231; Ts, Thioalkalivibrio sulfidophilus HLEbGR7; Tv-K90, Thioalkalivibrio sp. K90mix; Va, Vibrio alginolyticus 12G01; Vc, Vibrio cholera O1; Vp, Vibrio parahaemolyticus RIMD2210633. The sequences of Thioalkalivibrio sp. K90mix are shown in bold type. The bar indicates 20% sequence difference. Numbers on the branches indicate percentage bootstrap values from 1000 iterations; only those values are shown that distinguish the different flagellar motors. The CheW sequence of Thioalkalivibrio sp. K90mix (TK90_1183) was used as an outgroup, but was pruned from the tree.

Transposases and environmental stress

Comparative analysis of the genomes of “Thioalkalivibrio sulfidophilus” HL-EbGr7 [32] and Thioalkalivibrio sp. K90mix showed a greater abundance of genes encoding different transposases (i.e., COG2801, COG3328, COG3547) in the latter. Transposases are enzymes that can move specific sequences of DNA, known as transposons or transposable elements, within the genome. Krulwich [51] found that the genome of the alkaliphilic bacterium Bacillus halodurans C125 contained 112 transposase genes as compared to 10 in the genome of its closest non-alkaliphilic relative B. subtilis. She suggested that this might be one of the mechanisms of alkaliphilic adaptation at the genome level. Although strains HL-EbGr7 and K90mix are both obligately alkaliphilic, they differ in salt tolerance. HL-EbGr7 can tolerate only low (up to 1.5 M) salt concentrations, while K90mix can tolerate high (up to 4 M) salinities.

Capy et al. [52] mentioned that environmental stress might stimulate transposition and consequently increase the genetic variability, which can be beneficial for the adaptation to novel environmental conditions. Foti et al. [4] used rep-PCR [53] to study the genetic diversity within the genus Thioalkalivibrio and found a relatively high diversity of 56 genotypes among 85 strains that were isolated from different soda lakes in Africa and Asia. In addition, preliminary enrichment experiments with potassium carbonate instead of sodium carbonate and higher concentrations of chloride selected populations of high salt-tolerant Thioalkalivibrio strains with different rep-PCR patterns (unpublished results), which might be an indication that transposition might occur more frequently in strains with a wide range of salt tolerance.

Oxidative stress

Reactive oxygen species (ROS), such as superoxides (O2-) and hydrogen peroxidase (H2O2), are naturally produced at hypersaline conditions and are deleterious to cellular macromolecules. To protect themselves from this oxidative stress, Thioalkalivibrio sp. K90mix and “Thioalkalivibrio sulfidophilus” HL-EbGr7 have several defense mechanisms. Superoxides are converted to oxygen and hydrogen peroxide by the enzyme superoxide dismutase (TK90_0947, Tgr7_2463), while hydrogen peroxide is converted to oxygen by hydroperoxidase (TK90_0947, Tgr7_1107) or to H2O by the cytochrome C peroxidase (TK90_0812, Tgr7_2739). In addition, Thioalkalivibrio sp. K90mix produces high concentrations of a specific membrane-bound yellow pigment named ‘natronochrome’ [54]. The pigment has a high degree of unsaturation and might also play a role in the protection against reactive oxygen species (ROS). The gene(s) responsible for the synthesis of this anti-oxidant remains to be identified.

Osmotic stress

Thioalkalivibrio sp. K90mix is an extremely salt-tolerant bacterium. It can grow in saturated sodium and potassium carbonate and sodium sulfate brines containing up to 4 M Na+/K+ but, in contrast to halo-alkaliphiles, it is inhibited by high concentrations of chloride. So, a more proper term for such an extremophile would be an “extreme natronophile”. To withstand these extreme salinities, it synthesizes glycine betaine as the main compatible solute; the genome contains the genes for glycine sarcosine N-methyltransferase (TK90_0179) and sarcosine dimethylglycine methyltransferase (TK90_0180). In addition, the genome contains the gene for sucrose phosphate synthase (TK90_2312) to produce sucrose as a compatible solute. Genes for ectoine were not found.

References

  1. Sorokin DY, Kuenen JG. Haloalkaliphilic sulfur-oxidizing bacteria from soda lakes. FEMS Microbiol Rev 2005; 29:685–702. PubMed doi:10.1016/j.femsre.2004.10.005

    Article  CAS  PubMed  Google Scholar 

  2. Sorokin DY, Banciu H, Robertson LA, Kuenen JG. Haloalkaliphilic sulfur-oxidizing bacteria. In: The Prokaryotes. Volume 2: Ecophysiology and Biochemistry 2006; pp. 969–984. Dworkin, M., Falkow S, Rosenberg E, Schleifer KH, Stackebrandt E. (Ed’s). Springer, New York.

    Chapter  Google Scholar 

  3. Sorokin DY, van den Bosch PLF, Abbas B, Janssen AJH, Muyzer G. Microbiological analysis of the population of extremely haloalkaliphilic sulfur-oxidizing bacteria dominating in lab-scale sulfide-removing bioreactors. Appl Microbiol Biotechnol 2008; 80:965–975. PubMed doi:10.1007/s00253-008-1598-8

    Article  CAS  PubMed  Google Scholar 

  4. Foti M, Ma S, Sorokin DY, Rademaker JLW, Kuenen GJ, Muyzer G. Genetic diversity and bio-geography of haloalkaliphilic sulfur-oxidizing bacteria beloning to the genus Thioalkalivibrio. FEMS Microbiol Ecol 2006; 56:95–101. PubMed doi:10.1111/j.1574-6941.2006.00068.x

    Article  CAS  PubMed  Google Scholar 

  5. Sorokin DY, Kuenen JG, Jetten M. Denitrification at extremely alkaline conditions in obligately autotrophic alkaliphilic sulfur-oxidizing bacterium “Thioalkalivibrio denitrificans”. Arch Microbiol 2001; 175:94–101. PubMed doi:10.1007/s002030000210

    Article  CAS  PubMed  Google Scholar 

  6. Sorokin DY, Antipov AN, Kuenen JG. Complete denitrification in a coculture of haloalkaliphilic sulfur-oxidizing bacteria from a soda lake. Arch Microbiol 2003; 180:127–133. PubMed doi:10.1007/s00203-003-0567-y

    Article  CAS  PubMed  Google Scholar 

  7. Sorokin DY, Tourova TP, Lysenko AM, Kuenen JG. Microbial thiocyanate utilization under high alkaline conditions. Appl Environ Microbiol 2001; 67:528–538. PubMed doi:10.1128/AEM.67.2.528-538.2001

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  8. Sorokin DY, Tourova TP, Antipov AN, Muyzer G, Kuenen JG. Anaerobic growth of the haloalkaliphilic denitrifying sulfur-oxidizing bacterium Thialkalivibrio thiocyanodenitrificans sp. nov. with thiocyanate. Microbiology 2004; 150:2435–2442. PubMed doi:10.1099/mic.0.27015-0

    Article  CAS  PubMed  Google Scholar 

  9. Janssen AJ, Lens PN, Stams AJ, Plugge CM, Sorokin DY, Muyzer G, Dijkman H, Van Zessen E, Luimes P, Buisman CJ. Application of bacteria involved in the biological sulfur cycle for paper mill effluent purification. Sci Total Environ 2009; 407:1333–1343. PubMed doi:10.1016/j.scitotenv.2008.09.054

    Article  CAS  PubMed  Google Scholar 

  10. van den Bosch PLF, Sorokin DY, Buisman CJN, Janssen AJH. The effect of pH on thiosulfate formation in a new biotechnological process for the removal of hydrogen sulfide from gas streams. Environ Sci Technol 2008; 42:2637–2642. PubMed doi:10.1021/es7024438

    Article  PubMed  Google Scholar 

  11. 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 

  12. 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 

  13. Garrity GM, Bell JA, Lilburn T. Phylum XIV. Proteobacteria phyl. nov. In: Garrity GM, Brenner DJ, Krieg NR, Staley JT (eds), Bergey’s Manual of Systematic Bacteriology, Second Edition, Volume 2, Part B, Springer, New York, 2005, p. 1.

    Chapter  Google Scholar 

  14. List Editor. Validation of publication of new names and new combinations previously effectively published outside the IJSEM. List No. 106. Int J Syst Evol Microbiol 2005; 55:2235–2238. doi:10.1099/ijs.0.64108-0

  15. Garrity GM, Bell JA, Lilburn T. Class III. Gamma-proteobacteria class. nov. In: Garrity GM, Brenner DJ, Krieg NR, Staley JT (eds), Bergey’s Manual of Systematic Bacteriology, Second Edition, Volume 2, Part B, Springer, New York, 2005, p. 1.

    Chapter  Google Scholar 

  16. Imhoff J. Order I. Chromatiales ord. nov. In: Garrity GM, Brenner DJ, Krieg NR, Staley JT (eds), Bergey’s Manual of Systematic Bacteriology, Second Edition, Volume 2, Part B, Springer, New York, 2005, p. 1–3.

    Chapter  Google Scholar 

  17. Imhoff JF. Reassignment of the genus Ectothiorhodospira Pelsh 1936 to a new family, Ectothiorhodospiraceae fam. nov., and emended description of the Chromatiaceae Bavendamm 1924. Int J Syst Bacteriol 1984; 34:338–339. doi:10.1099/00207713-34-3-338

    Article  Google Scholar 

  18. Sorokin DY, Lysenko AM, Mityushina LL, Tourova TP, Jones BE, Rainey FA, Robertson LA, Kuenen GJ. Thioalkalimicrobium aerophilum gen. nov., sp. nov. and Thioalkalimicrobium sibericum sp. nov., and Thioalkalivibrio versutus gen. nov., sp. nov., Thioalkalivibrio nitratis sp.nov., novel and Thioalkalivibrio denitrificancs sp. nov., novel obligately alkaliphilic and obligately chemolithoautotrophic sulfur-oxidizing bacteria from soda lakes. Int J Syst Evol Microbiol 2001; 51:565–580. PubMed

    Article  CAS  PubMed  Google Scholar 

  19. Banciu H, Sorokin DY, Galinski EA, Muyzer G, Kleerebezem R, Kuenen JG. Thialkalivibrio halophilus sp. nov., a novel obligately chemolithoautotrophic, facultatively alkaliphilic, and extremely salt-tolerant, sulfur-oxidizing bacterium from a hypersaline alkaline lake. Extremophiles 2004; 8:325–334. PubMed doi:10.1007/s00792-004-0391-6

    CAS  PubMed  Google Scholar 

  20. List Editor. Notification that new names and new combinations have appeared in volume 51, part 2, of the IJSEM. Int J Syst Evol Microbiol 2001; 51:795–796. PubMed doi:10.1099/00207713-51-3-795

  21. De Vos P, Trüper HG, Tindall BJ. Judicial Commission of the International Committee on Systematics of Prokaryotes Xth International (IUMS) Congress of Bacteriology and Applied Microbiology. Minutes of the meetings, 28, 29 and 31 July and 1 August 2002, Paris, France. Int J Syst Evol Microbiol 2005; 55:525–532. doi:10.1099/ijs.0.63585-0

    Article  Google Scholar 

  22. Classification of Bacteria and Archaea in risk groups. www.baua.de TRBA 466.

  23. http://www.iTouchMap.com

  24. Pruesse E, Quast C, Knittel K, Fuchs B, Ludwig W, Peplies J, Glöckner FO. SILVA: a comprehensive online resource for quality checked and aligned ribosomal RNA sequence data compatible with ARB. Nucleic Acids Res 2007; 35:7188–7196. PubMed doi:10.1093/nar/gkm864

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  25. Ludwig W, Strunk O, Westram R, Richter L, Meier H. Yadhukumar, Buchner A, Lai T, Steppi S, Jobb G. ARB: a software environment for sequence data. Nucleic Acids Res 2004; 32:1363–1371. PubMed doi:10.1093/nar/gkh293

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  26. Liolios K, Chen IM, Mavromatis K, Tavernarakis N, Hugenholtz P, Markowitz VM, Kyrpides NC. The Genomes OnLine 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 

  27. DOE Joint Genome Institute. http://www.jgi.doe.gov/

  28. Lapidus A, LaButti K, Foster B, Lowry S, Trong S, Goltsman E. POLISHER: An effective tool for using ultra short reads in microbial genome assembly and finishing. AGBT, Marco Island, FL, 2008.

    Google Scholar 

  29. Hyatt D, Chen GL, LoCascio 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 

  30. Pati A, Ivanova NN, Mikhailova N, Ovchinnikova G, Hooper SD, Lykidis A, Kyrpides NC. GenePRIMP: 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 

  31. Markowitz VM, Szeto E, Palaniappan K, Grechkin Y, Chu K, Chen IM, Dubchak I, Anderson I, Lykidis A, Mavromatis K, et al. The integrated microbial genomes (IMG) system in 2007: data content and analysis tools extensions. Nucleic Acids Res 2008; 36:D528–D533. PubMed doi:10.1093/nar/gkm846

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  32. Muyzer G, Sorokin DY, Mavromatis K, Lapidus A, Clum A, Ivanova N, Pati A, d’Haeseleer P, Woyke T, Kyrpides NC. Complete genome sequence of “Thioalkalivibrio sulfidophilus” HL-EbGR7. Stand Genomic Sci 2011; 4:23–35. PubMed doi:10.4056/sigs.1483693

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  33. Shibata M, Katoh H, Sonoda M, Ohkawat H, Shimoyama M, Fukuzawa H, Kaplan A, Ogawa T. Genes essential to sodium-dependent bicarbonate transport in cyanobacteria. J Biol Chem 2002; 277:18658–18664. PubMed doi:10.1074/jbc.M112468200

    Article  CAS  PubMed  Google Scholar 

  34. Gibson J, Pfennig N, Waterbury JB. Chloroherpeton thalassium gen. nov. et spec. nov., a non-filamentous, flexing and gliding green sulfur bacterium. Arch Microbiol 1984; 138:96–101. PubMed doi:10.1007/BF00413007

    Article  CAS  PubMed  Google Scholar 

  35. Frigaard N-U, Dahl C. Sulfur metabolism in phototrophic sulfur bacteria. Advances in Microbial Physiology, 2009; 54: 1–3–200.

    Google Scholar 

  36. Hanson TE, Tabita FR. A ribulose-1,5-biphosphate carboxylase/oxygenase (RubisCO)-like protein from Chlorobium tepidum that is involved with sulfur metabolism and the response to oxidative stress. Proc Natl Acad Sci USA 2001; 98:4397–4402. PubMed doi:10.1073/pnas.081610398

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  37. Rohwerder T, Sand W. The sulfane sulfur of persulfides is the actual substrate of the sulfur-oxidizing enzymes from Acidithiobacillus and Acidiphillum spp. Microbiology 2003; 149:1699–1710. PubMed doi:10.1099/mic.0.26212-0

    Article  CAS  PubMed  Google Scholar 

  38. Hedderich R, Hamann N, Bennati M. Heterodi-sulfide reductase from methanogenic archaea: a new catalytic role of iron-sulfur cluster. Biol Chem 2005; 386:961–970. PubMed doi:10.1515/BC.2005.112

    Article  CAS  PubMed  Google Scholar 

  39. Mander GJ, Pierik AJ, Huber H, Hedderich R. Two distinct heterodisulfide reductase-like enzymes in the sulfate-reducing archaeon Archaeoglobus profundus. Eur J Biochem 2004; 271:1106–1116. PubMed doi:10.1111/j.1432-1033.2004.04013.x

    Article  CAS  PubMed  Google Scholar 

  40. Quatrini R, Appia-Ayme C, Denis Y, Jedlicki E, Holmes DS, Bonnefoy V. Extending the model for iron and sulfur oxidation in the extreme acidophile Acidithiobacillus ferrooxidans. BMC Genomics 2009; 10:394. PubMed doi:10.1186/1471-2164-10-394

    Article  PubMed Central  PubMed  Google Scholar 

  41. Mangold S, Valdés J, Holmes DS, Dopson M. Sulfur metabolism in the extreme acidophile Acidithiobacillus caldus. Front Microbiol 2011; 2:1–13. PubMed

    Article  Google Scholar 

  42. Kappler U. Bacterial sulfite-oxidizing enzymes. Biochim Biophys Acta 2011; 1807: 1–10.

    Article  CAS  PubMed  Google Scholar 

  43. Oren A. Thermodynamic limits to microbial life at high salt concentrations. [PubMed]. Environ Microbiol 2011; 11:1908–1923. PubMed

    Article  Google Scholar 

  44. Biegel E, Müller V. Bacterial Na+-translocating ferridoxins: NAD+ oxidoreductase. Proc Natl Acad Sci USA 2010; 107:18138–18142. PubMed doi:10.1073/pnas.1010318107

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  45. Padan E, Bibi E, Ito M, Krulwich TA. Alkaline pH homeostasis in bacteria: New insights. Biochim Biophys Acta 2005; 1717:67–88. PubMed doi:10.1016/j.bbamem.2005.09.010

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  46. Price GD, Woodger FJ, Badger MR, Howitt SM, Tucker L. Identification of a SulP-type bicarbonate transporter in marine cyanobacteria. Proc Natl Acad Sci USA 2004; 101:18228–18233. PubMed doi:10.1073/pnas.0405211101

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  47. Corratgé-Faillie C, Jabnoune M, Zimmerman S, Very AA, Fizames C, Sentenac H. Potassium and sodium transport in non-animals cells: the Trk/Ktr/HKT transporter system. Cell Mol Life Sci 2010; 67:2511–2532. PubMed doi:10.1007/s00018-010-0317-7

    Article  PubMed  Google Scholar 

  48. Porter SL, Wadhams GH, Armitage JP. Signal processing in complex Chemotaxis pathways. Nat Rev Microbiol 2011; 9:153–165. PubMed doi:10.1038/nrmicro2505

    Article  CAS  PubMed  Google Scholar 

  49. Pallen MJ, Matzke NJ. From the origin of species to the origin of bacterial flagella. Nat Rev Microbiol 2006; 4:784–790. PubMed doi:10.1038/nrmicro1493

    Article  CAS  PubMed  Google Scholar 

  50. Krulwich TA. Bioenergetic adaptations that support alkaliphily. 2007. In: Physiology and biochemistry of extremophiles. Eds. Gerday C, Glansdorff N. ASM Press, Washington, D.C. pp. 311–329.

    Chapter  Google Scholar 

  51. Krulwich TA. Alkaliphily. 2003. In: Extremophiles (Life under extreme conditions) [Eds. Gerday C, Glansdorff N], in Encyclopedia of Life Support Systems (EOLSS). UNESCO-EOLSS.

  52. Capy P, Gasperi G, Biémont C, Bazin C. Stress and transposable element: co-evolution or useful parasites? Heredity 2000; 85:101–106. PubMed doi:10.1046/j.1365-2540.2000.00751.x

    Article  CAS  PubMed  Google Scholar 

  53. Versalovic J, de Bruijn FJ, Lupski JR. Genomic fingerprinting of bacteria using repetitive sequence based PCR 9rep-PCR). Meth Cell Mol Biol 1994; 5:25–40.

    CAS  Google Scholar 

  54. Takaichi S, Maoka T, Akimoto N, Sorokin DY, Banciu H, Kuenen JG. Two novel yellow pigments natronochrome and chloronatronochrome from the natrono(alkali)philic sulfur-oxidizing bacterium Thialkalivibrio versutus ALJ 15. Tetrahedron Lett 2004; 45:8303–8305. doi:10.1016/j.tetlet.2004.09.073

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was performed under the auspices of the US Department of Energy Office of Science, Biological and Environmental Research Program, and by the University of California, Lawrence Berkeley National Laboratory under contract No. DE-AC02-05CH11231, Lawrence Livermore National Laboratory under Contract No. DE-AC52-07NA27344, and Los Alamos National Laboratory under contract No. DE-AC02-06NA25396, UT-Battelle and Oak Ridge National Laboratory under contract DE-AC05-00OR22725. Dimitry Sorokin was supported financially by RFBR grant 10-04-00152.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gerard Muyzer.

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

Muyzer, G., Sorokin, D.Y., Mavromatis, K. et al. Complete genome sequence of Thioalkalivibrio sp. K90mix. Stand in Genomic Sci 5, 341–355 (2011). https://doi.org/10.4056/sigs.2315092

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.4056/sigs.2315092

Keywords