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

High-quality-draft genome sequence of the fermenting bacterium Anaerobium acetethylicum type strain GluBS11T (DSM 29698)

Abstract

Anaerobium acetethylicum strain GluBS11T belongs to the family Lachnospiraceae within the order Clostridiales. It is a Gram-positive, non-motile and strictly anaerobic bacterium isolated from biogas slurry that was originally enriched with gluconate as carbon source (Patil, et al., Int J Syst Evol Microbiol 65:3289-3296, 2015). Here we describe the draft genome sequence of strain GluBS11T and provide a detailed insight into its physiological and metabolic features. The draft genome sequence generated 4,609,043 bp, distributed among 105 scaffolds assembled using the SPAdes genome assembler method. It comprises in total 4,132 genes, of which 4,008 were predicted to be protein coding genes, 124 RNA genes and 867 pseudogenes. The G + C content was 43.51 mol %. The annotated genome of strain GluBS11T contains putative genes coding for the pentose phosphate pathway, the Embden-Meyerhoff-Parnas pathway, the Entner-Doudoroff pathway and the tricarboxylic acid cycle. The genome revealed the presence of most of the necessary genes required for the fermentation of glucose and gluconate to acetate, ethanol, and hydrogen gas. However, a candidate gene for production of formate was not identified.

Introduction

Strain GluBS11T (= DSM 29698) is the type strain of the newly described species Anaerobium acetethylicum [1]. The genus Anaerobium belongs to the family Lachnospiraceae [2] within the class Clostridia [3] of the order Clostridiales [4] that is largely synonymous with Clostridium rRNA cluster XIVa [5, 6]. Members of the family Lachnospiraceae have been isolated from diverse habitats, but are mainly constituents of mammalian intestinal microbiota, especially from ruminants [7] and humans [8]. They are strictly anaerobic and primarily non-spore forming [9], and ferment polysaccharides to short-chain fatty acids such as acetate and propionate as fermentation products [10], e.g., Eubacterium rectale ATCC 33656 T, Eubacterium ventriosum ATCC 27560 T, Coprococcus sp. and Roseburia sp. [11, 12]. The family Lachnospiraceae as currently described in the National Center for Biotechnology Information homepage comprises 41 named genera and several unclassified isolates, of which a total of 143 draft or complete genome sequences are available. Strain GluBS11T was isolated due to its ability to ferment gluconate, and the species epithet ‘acetethylicum’ refers to its main fermentation products acetate and ethanol during gluconate fermentation [1]. Within the diverse family of Lachnospiraceae , strain GluBS11T is phylogenetically closely related to the type strains of C. herbivorans strain 54408 [94% 16S rRNA sequence similarity); [13], C. populeti ATCC 35295 T (93.3% similarity); [14], Eubacterium uniforme ATCC 35992 T (92.4% similarity), and C. polysaccharolyticum ATCC 33142 T (91.5% similarity); [15, 16]. Of these, all strains were reported to ferment sugars mainly to butyrate plus formate, acetate, ethanol or lactate, except E. uniforme , which does not produce butyrate. Similar to E. uniforme ATCC 35992 T, strain GluBS11T does not produce butyrate during the fermentation of sugars or glycerol [1, 17]. Moreover, none of the above strains except for strain GluBS11T was tested for fermentation of gluconate.

The most prominent feature of A. acetethylicum strain GluBS11T is its ability to ferment sugars (including oxidized sugar such as gluconate) and glycerol mainly to acetate, ethanol, hydrogen, and formate [1, 17]. Therefore, we selected strain GluBS11T as a candidate for studying its potential to ferment gluconate or glycerol. Moreover, most of the described bacterial glycerol fermentations lead to 1,3-propanediol [18] and other undesired products such as butyrate or 2,3-butanediol. In contrast to this, strain GluBS11T ferments glycerol mainly to ethanol and hydrogen gas as well as negligible amounts of acetate and formate [17]. Here we present the summary of the taxonomic classification and the features of A. acetethylicum strain GluBS11T together with the description of the genome sequencing and annotation. Emphasis is given on understanding the central metabolism and fermentation pathways. The putative enzymes involved in the fermentation of gluconate, glucose, and glycerol were also reconstructed from the genomic data.

Organism information

Classification and features

A. acetethylicum strain GluBS11T is a member of the family Lachnospiraceae in the phylum Firmicutes [19]. Cells were strictly anaerobic, non-motile and stained Gram-positive [1]. Fig. 1a shows the ultrathin trans-section of a rod-shaped cell and Fig. 1b shows details of the Gram-positive membrane structure. For transmission electron microscopy, fixation of bacterial cells was done with glutardialdehyde and osmium tetroxide followed by staining with uranylacetate. Samples were dehydrated in a graded ethanol series, embedded in Spurr resin and viewed in a Zeiss 912 Omega transmission electron microscope (Oberkochen, Germany) at 80 kV. Classification and general features are summarized in Table 1. Strain GluBS11T ferments various substrates including glucose, lactose, sucrose, fructose, maltose, xylose, galactose, melibiose, melezitose, gluconate, mannitol, erythritol, glycerol and esculin, and mainly produces acetate, ethanol, hydrogen and formate as fermentation end products [1]. Although strain GluBS11T was tested negative for catalase and peroxidase [1]. A gene coding for a putative catalase-peroxidase (IMG gene locus tag Ga0116910_10254) was identified in the draft genome. Besides this, strain GluBS11T contains putative genes coding for thioredoxin reductase (Ga0116910_100846) and thioredoxin (Ga0116910_100229), and no gene coding for superoxide dismutase was identified in the genome. Strain GluBS11T was tested positive for fermentation (API Rapid 32A reactions) of α-galactosidase, α-glucosidase and β-glucosidase [1]. The genome-based analysis identified genes coding for a putative β-galactosidase (Ga0116910_1001515 and Ga0116910_100295), a β-glucosidase (Ga0116910_100187 and Ga0116910_100196) and α-galactosidase (Ga0116910_10579, Ga0116910_100577 and Ga0116910_102538), respectively.

Fig. 1
figure 1

Transmission electron micrograph of A. acetethylicum strain GluBS11T cells grown with gluconate. a Ultrathin trans-section of cell; b details of the Gram-positive membrane structure (white arrows)

Table 1 Classification and general features of Anaerobium acetethylicum strain GluBS11T according to the MIGS recommendations [53]

BLAST search results of the partial 16S rRNA gene sequence of A. acetethylicum strain GluBS11T (KP233894) revealed closest sequence similarities with the uncultured Lachnospiraceae bacterium strain UY038 (94% similarity; HM099641) that was isolated from an oral sample, C. populeti ATCC 35295 T (94%; X71853) and Robinsoniella sp. MCWD5 (94%; KU886099). The draft genome sequence of A. acetethylicum GluBS11T has one full-length 16S rRNA gene (1,536 bp; locus tag Ga0116910_1073) that was compared with the partial 16S rRNA gene sequence (1,402 bp; KP233894) from the original species description [1]. Sequence alignment indicated, that both 16S rRNA sequences were about 99% identical and the complete 16S rRNA gene sequence differs from the partial 16S rRNA gene sequence by the presence of an additional stretch of 45 bp long nucleotide sequence at the beginning, 5 gaps (53-55, 65 and 68 positions), and 9 base change at position 51 (T-A), 96 (G-A), 104 (A-T), 1,008 (T-A), 1,423 (A-T), 1,434 (A-G), 1,435 (T-G), 1,442 (A-C) and 1443 (T-C), followed by an additional long stretch of a 83 bp nucleotide sequence at the end. Figure 2 shows the current phylogenetic position of A. acetethylicum strain GluBS11T in a phylogenetic tree constructed in MEGA 7 [20] using the Minimum Evolution method [21], and the evolutionary distances were computed using the Jukes-Cantor method [22] and the Neighbor-Joining algorithm [23].

Fig. 2
figure 2

Phylogenetic tree constructed using MEGA 7 [20] showing the current position of the A. acetethylicum strain GluBS11T with respect to the selected members from the order Clostridiales. The evolutionary distances were computed using the Jukes-Cantor method [22] and are in the units of the number of base substitutions per site. The phylogenetic tree was searched using the Close-Neighbor-Interchange algorithm [59] at a search level of 1. All positions containing gaps and missing data were eliminated. There were a total of 1,300 positions in the final dataset. Numbers at the nodes indicates the bootstrap values from 1000 replicates [60] and accession numbers are given in parentheses. Bar indicates 2% estimated sequence divergence

Genome sequencing information

Genome project history

Strain GluBS11T was selected for genome sequencing because of its ability to ferment gluconate or glycerol mainly to acetate, ethanol, hydrogen and small amounts of formate. Genome sequencing was performed through the community science program as part of the “Genomic Encyclopedia of Bacterial and Archaeal Type Strains, Phase III: the genomes of soil and plant-associated and newly described type strains” [24, 25]. The draft genome of A. acetethylicum strain GluBS11T is listed in the Genomes OnLine Database under the GOLD project ID Gp0139288 [26], and the assembled and annotated high-quality permanent draft genome sequence is deposited in IMG under submission ID 88715 [27]. Whole genome shotgun sequencing project was also submitted to the Genbank/NCBI under the accession no., FMKA00000000 and consists of 105 contigs (FMKA01000001-FMKA01000105). Sequencing, finishing and annotation were performed by the Department of Energy, Joint Genome Institute using state-of-the-art sequencing technology [28]. A summary of the project information is shown in Table 2.

Table 2 Project information

Growth conditions and genomic DNA preparation

A. acetethylicum strain GluBS11T was cultivated in anoxic mineral medium supplemented with 10 mM gluconate as growth substrate at 30 °C for three days until OD600nm 1.0 was reached. Genomic DNA was isolated from the cell pellet obtained from about 500 ml of grown culture using a CTAB-based method [29] with slight modifications [30]. After RNase treatment, the purity and quality of the genomic DNA preparation were assessed by DNA absorption at 260 nm and size by agarose gel electrophoresis (1% w/v; Additional file 1: Figure S1). The concentration of the isolated genomic DNA was 2.4 μg μl-1 (A260/280 = 2.03 and A260/230 = 2.47). Finally, the DNA was used to amplify the 16S rRNA gene to confirm the identity of genomic DNA by comparing with the described partial 16S rRNA gene sequence (KP233894) of A. acetethylicum strain GluBS11T. The pure and high-quality genomic DNA was shipped to the DOE, JGI for genome sequencing.

Genome sequencing and assembly

The draft genome sequencing was performed at the DOE, JGI using the Illumina technology [31]. An Illumina 300 bp insert standard shotgun library was constructed and sequenced using the Illumina HiSeq-2500 1 TB platform, which generated 11,508,336 reads totaling 1,726.3 Mbp. All details on library construction and sequencing performed at the JGI can be found on the website. All raw Illumina sequence data were filtered using BBDuk [32], which removes known Illumina artifacts and PhiX. Reads with more than one “N” or with quality scores (before trimming) averaging less than 8 or reads shorter than 51 bp (after trimming) were discarded. Remaining reads were mapped to masked versions of human, cat and dog references using BBMap [32] and discarded if the identity exceeded 95%. Sequence masking was performed with BBMask [32]. The following steps were performed for assembly: (1) artifact filtered Illumina reads were assembled using the SPAdes genome assembler (version 3.6.2); [33], (2) assembly contigs were discarded if their length was <1 kbp. Parameters for the SPAdes assembly were -cov-cutoff auto -phred-offset 33 -t 8 -m 40 -careful -k 255,595 -12. The final draft assembly contained 108 contigs in 105 scaffolds, totaling 4.609 Mbp in size, and was based on 1,500.0 Mbp of Illumina data with a mapped coverage of 336.0X.

Genome annotation

Genes were identified with Prodigal [34] using standard microbial genome annotation pipeline [35]. The predicted CDSs were translated and used to search the NCBI non-redundant database, UniProt, TIGRFam, Pfam, KEGG, COG, and InterPro databases. The tRNAScanSE tool [36] was used to find tRNA genes, whereas rRNA genes were found by searches against models of the rRNA genes built from SILVA [37]. Other non-coding RNAs such as 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 [38]. Additional gene prediction analysis and manual functional annotation (IMG taxon ID 2675903067) were performed within the Integrated Microbial Genomes-Expert Review platform [39] developed by the JGI, Walnut Creek, CA, USA.

Genome properties

The draft genome sequence of A. acetethylicum strain GluBS11T was based on an assembly of 105 DNA scaffolds (108 contigs) amounting to 4,609,043 (4.6 Mb) nucleotide base pairs with a calculated G + C content of 43.51 mol % (Table 3). Of the total of predicted CDSs of 4,132 genes (100%), 4,008 were assigned to protein-coding genes, of which 2,640 were assigned to COGs (63.89%), and the rest of 124 were assigned to RNA genes (3.0%). The majority of protein-coding genes (3,141 genes or 76.02%) were assigned to putative functions whilst the remaining genes were annotated as hypothetical proteins of unknown function. The draft genome properties, the statistics and the distribution of genes into COGs functional categories are summarized in Tables 3 and 4. The draft genome comparison of A. acetethylicum strain GluBS11T using the BLASTn revealed top hits with the genomes of C. nexile DSM 1787 T (85% identity; NZ_DS995342.4), Anaerostipes hadrus DSM 3319 T (85%; NZ_KB290653.1), Acetonema longum DSM 6540 T (84%; NZ_AFGF01000168.1), Anaerostipes caccae DSM 14662 T (83%; NZ_DS499733.1), Blautia hansenii DSM 20583 T (83%; NZ_GG698589.1), and a ruman-associated strain, Ruminococcus torques ATCC 27756 T (82%; NZ_DS264349.1), and C. phytofermentans ATCC 700394 T (74%), respectively.

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

Insights from the genome sequence

General metabolic features

The draft genome of strain GluBS11T was further examined to understand the organism’s physiology and fermentation metabolism. The draft genome encodes most of the key enzymes of the pentose phosphate pathway, Embden-Meyerhoff-Parnas pathway, Entner-Doudoroff pathway and tricarboxylic acid cycle (Additional file 2: Table S1). Thus, strain GluBS11T is very likely to use these pathways for its central metabolism and biosynthesis. Besides this, the genome also contains the genes coding for putative enzymes of anaplerotic pathways, such as oxaloacetate decarboxylase (α-subunit, Ga0116910_1001318 and β-subunit, Ga0116910_1001319), pyruvate kinase (Ga0116910_1001611), fructose-1,6-bisphosphatase (Ga0116910_1001181 and 10346), phosphoenolpyruvate carboxykinase (Ga0116910_1001300) and pyruvate carboxylase β-subunit (Ga0116910_101716). Genes for biosynthesis of amino acids and most co-factors were also present (Additional file 2: Table S1).

Although cells of strain GluBS11T are non-motile [1], the genome possesses genes that are predicted to encode flagellum components (Ga0116910_1001565, Ga0116910_1002133- Ga0116910_1002135, Ga0116910_100329, Ga0116910_1002133- Ga0116910_1002135) such as flagellar protein FliO/FliZ, flagellar motor switch protein FliN/FliY/FliM, flagellar FliL protein, and pilus assembly-protein (Flp/PilA), which are located in a single gene cluster (locus tag Ga0116910_100336 to Ga0116910_100363), including the chemotaxis protein (MotB/A). The draft genome also contains genes predicted to encode seven universal stress proteins of the UspA family (gene loci Ga0116910_103114, 1003225, 10025, 10028, 10027, 104111 and 100540), 2 heat-shock proteins such as GrpE (Ga0116910_10476 and 100386), one heat-inducible transcriptional repressor (Ga0116910_100387), and six cold-shock proteins of the CspA family (Ga0116910_10067, 1002200, 1001175, 1004187, 1005160 and 1002190). Also, a DNA-directed RNA polymerase with sigma-70/32 factor (ECF family) and a heat-inducible transcriptional repressor (HrcA) along with the RNA polymerase sigma factor for flagellar operon FliA were detected in the draft genome.

Clustered regularly interspaced short palindromic repeats are segments of prokaryotic DNA containing short repetitions of base sequences followed by a short segment of ‘spacer DNA’ that function as a defense system against the introduction of foreign genetic materials (e.g., phage infection, plasmid or horizontal gene transfer). CRISPRs were found in approximately 40% of all sequenced bacterial genomes [40]. Genome analysis of strain GluBS11T suggests that the genome does not contain CRISPR regions, although the genome of the phylogenetically closely related strain C. populeti ATCC 35295 T contains two gene coding for CRISPR-associated proteins (cas9 family protein; Ga0056054_02523 and Ga0056054_00025).

Transporters

Transporters enable bacteria to accumulate required nutrients and also contribute for excretion of unwanted metabolic products. They also help to maintain the osmotic balance and the cytoplasmic pH by transporting H+ and various salts. Genome analysis of strain GluBS11T identified various membrane transporters including the ABC solute transporters (ATP-dependent) that could take part in the transport of various substrates such as ions, vitamins, sugars, amino acids, and metabolites (Additional file 3: Table S2). Most of these identified transporters belong to diverse transporter families such as the amino acid/polyamine antiporter family, the drug/metabolite transporter superfamily, and the major facilitator superfamily that is used for transport of a diverse set of small solutes in response to chemiosmotic ion gradients [41]. The draft genome sequence also contains several genes coding for proton symporters (Additional file 3: Table S2). Thus, strain GluBS11T could generate a proton gradient using FoF1-type ATP synthase in reverse direction [42, 43].

Metabolic pathways for glucose, gluconate and glycerol utilization

Strain GluBS11T ferments sugars, e.g., glucose and gluconate or glycerol mainly to ethanol and hydrogen, including the production of acetate and small amounts of formate as fermentation end products [1, 17]. In the present study, a metabolic network for the utilization of glucose and gluconate including glycerol was constructed based on the genome as shown in Fig. 3, from the genome annotation provided by the IMG-ER. To determine which pathway was utilized for glycerol fermentation, a recent study by Patil et al., [17] provided insight into glycerol fermentation of strain GluBS11T using biochemical and proteomic approaches. There are three possible alternatives for gluconate metabolism: first, the phosphorylation to gluconate 6-phosphate (the Entner-Doudoroff pathway), second, the reduction to glucose or lastly, the dehydration to 2-keto-3-deoxy-gluconate, a modified Entner-Doudoroff pathway [44]. In the last four decades, several studies reported that gluconate fermentation by numerous anaerobic bacteria, e.g., Clostridium aceticum DSM 1496 T [45] or E. coli ML30 (DSM 1328 T); [46] proceeds through a modified Entner-Doudoroff pathway.

Fig. 3
figure 3

Metabolic network of glucose and gluconate, including glycerol [17] metabolism by A. acetethylicum strain GluBS11T reconstructed from the IMG annotated draft genome sequence. Numbers adjacent to arrows represent putative enzymes. 1) 2-keto-3-deoxphosphogluconate aldolase (locus tag, Ga0116910_101517); 2) glycerol dehydrogenase (Ga0116910_101526 and 101551); 3) dihydroxyacetone kinase (Ga0116910_ 1001186, 1001188, 101527, 101552 and 101085); 4) triosephosphate isomerase (Ga0116910_ 1001390, 102914, 101435 and 101134); 5) phosphotransferase system (PTS; Ga0116910_100991 and Ga0116910_100370); 6) phosphogluconomutase (Ga0116910_ 1007105, 10644, 1002181 and 10031112); 7) phosphofructokinase (Ga0116910_100239); 8) fructose 1, 6-bisphosphate aldolase (Ga0116910_100167); 9) glyceraldehyde 3-phosphate dehydrogenase (Ga0116910_1001391); 10) phosphoglycerate kinase (Ga0116910_1001391); 11) phosphoglycerate mutase (Ga0116910_1001389 and Ga0116910_103027); 12) enolase (Ga0116910_1001503); 13) pyruvate kinase (Ga0116910_1004153); 14) pyruvate ferredoxin oxidoreductase (Ga0116910_103224 and Ga0116910_101718); 15) phosphoacetyl transferase (Ga0116910_1001587); 16) acetate kinase (Ga0116910_1001586); 17) CoA-dependent acetaldehyde dehydrogenase (Ga0116910_1004188); 18) alcohol dehydrogenase (Ga0116910_101528 and 101313); 19) iron-only hydrogenases (Ga0116910_100545, Ga0116910_1001473 and Ga0116910_100543); 20) NADP-reducing hydrogenases (Ga0116910_1001466,Ga0116910_1001467, Ga0116910_1001468, Ga0116910_1001470) and 21) putative pyruvate carboxylase (Ga0116910_101716)

The genome annotation predicted the presence of four gluconate:proton symporters (Gnt family) encoded by Ga0116910_10413, Ga0116910_10069, Ga0116910_100214 and Ga0116910_10418. In a previous study, it was shown that C. acetobutylicum ATCC 824 T takes up gluconate by gluconate:proton symporters (CA_C2835); [47] which showed amino acid sequence identity (24 to 41%) with the four predicted genes with highest identity (Ga0116910_10418; 42%). Thus, the product of the Ga0116910_10418 gene is the most likely candidate for uptake of gluconate. Based on the genome annotation, A. acetethylicum strain GluBS11T most likely uses the Entner-Doudoroff pathway for gluconate metabolism, through which gluconate is first phosphorylated to 6-phosphogluconate by gluconokinase (EC 2.7.1.12) followed by dehydration to 2-keto-3-deoxy-phosphogluconate by 6-phosphogluconate dehydratase (EC 4.2.1.12). Alternatively, gluconate could be first dehydrated (modified Entner-Doudoroff pathway) to 2-keto-3-deoxy gluconate by gluconate dehydratase (EC 4.2.1.39) followed by phosphorylation to KDPG by 2-keto-3-deoxygluconokinase (EC 2.7.1.45). KDPG would be further converted to pyruvate and glyceraldehyde 3-phosphate by KDPG aldolase (EC 4.1.2.14). The presence of a putative gene coding for KDPG aldolase (Ga0116910_101517) indicates that gluconate is most likely metabolized via KDPG. However, no putative genes coding for the initial enzymes that could convert gluconate to KDPG (according to two ways as mentioned above) was identified in the draft genome of strain GluBS11T. However, two putative genes were annotated as dihydroxy acid dehydratase/phosphogluconate dehydratase (Ga0116910_10068 and Ga0116910_101679) that could have this activity. The predicted dihydroxy acid dehydratase (EC 4.2.1.9) is possibly involved in the biosynthesis of amino acids (valine, isoleucine, and isoleucine). A similar observation was also reported for the gluconate-fermenting C. acetobutylicum ATCC 824 T, where the gene CA_C3170 was predicted to encode a 6-phosphogluconate dehydratase and BlastP analysis indicated that it is a dihydroxy acid dehydratase primarily involved in the synthesis of amino acids [47, 48]. BlastP search of amino acid sequence analysis of genes Ga0116910_10068 and Ga0116910_101679 showed more than 80% identity with the dihydroxy acid dehydratase of C. phytofermentans ATCC 700394 T (A9KL28) and Anaerostipes caccae DSM 14662 T, respectively, and showed only 40-60% identity with gene CA_C3170 of C. acetobutylicum ATCC 824 T. Therefore, genes Ga0116910_10068 and Ga0116910_101679 most likely encode a dihydroxy acid dehydratase that is involved in amino acid synthesis rather than in KDPG formation. Based on this information, gluconate degradation via the Entner-Doudoroff pathway involving gluconate phosphorylation to 6-phosphogluconate by gluconokinase (EC 2.7.1.12) followed by dehydration to KDPG by 6-phosphogluconate dehydratase (EC 4.2.1.12) can be ruled out. Furthermore, the presence of a putative gene coding for KDPG aldolase (Ga0116910_101517) indicates that gluconate is most likely metabolized via the modified Entner-Doudoroff pathway, which would be consistent with previous studies of the anaerobic gluconate metabolism [45, 47, 49]. Even though no genes coding for the gluconate dehydratase (EC 4.2.1.39) and KDG kinase (EC 2.7.1.178) required for initial activation of gluconate to KDPG were identified in the genome of strain GluBS11T.

While gluconate is predicted to be metabolized via the modified Entner-Doudoroff pathway, glucose could be metabolized through glycolysis. For uptake of glucose, strain GluBS11T most likely uses a phosphotransferase system (PTS) which couples glucose import to its phosphorylation with phosphoenolpyruvate, yielding glucose-6-phosphate and pyruvate [47]. Genes Ga0116910_100991 and Ga0116910_100370 are predicted to encode PTS proteins which are most likely involved in glucose transport in strain GluBS11T. Thus, genome analysis suggests that glucose is most probably metabolized through glycolysis via glucose 6-phosphate by glucose-6-phosphate isomerase (Ga0116910_1004120 and Ga0116910_10539), 6-phosphofructokinase (Ga0116910_103531, Ga0116910_100239, Ga0116910_102039 and Ga0116910_101135), and fructose-bisphosphate aldolase (Ga0116910_101128 and Ga0116910_102024) to glyceraldehyde 3-phosphate. In the glycolysis pathway, glyceraldehyde 3-phosphate is further metabolized through the lower part of glycolysis to ethanol, acetate, hydrogen, and formate. During gluconate fermentation, KDPG aldolase would then convert KDPG to glyceraldehyde-3-phosphate and pyruvate, and only glyceraldehyde-3-phosphate passes through the lower glycolysis pathway.

Previous studies with other bacteria reported that gluconate fermentation mainly yielded acetate and butyrate as fermentation products [45, 47, 49]. Although, the draft genome of strain GluBS11T contains genes predicted to code for a putative butyrate kinase (Ga0116910_101723 and Ga0116910_102110), gluconate, glucose or glycerol fermentation by strain GluBS11T does not produce butyrate [1, 17]. The pathways were easily constructed based on the genome analysis and genes for acetate metabolism, e.g., acetate kinase (Ga0116910_103636, Ga0116910_1001586 and Ga0116910_104214), ethanol metabolism, e.g., alcohol hydrogenase (Ga0116910_101528, Ga0116910_102038, Ga0116910_102215, Ga0116910_1004154 and Ga0116910_102016), and hydrogen metabolism, e.g., putative iron-only hydrogenases and subunits coding for an NADP+-reducing hydrogenase (Ga0116910_1001473, Ga0116910_1001466, Ga0116910_1001467, Ga0116910_1001468, Ga0116910_1001470, Ga0116910_100545 and Ga0116910_1001473). No candidate gene was found to code for a putative formate-producing formate dehydrogenase in the draft genome of strain GluBS11T even though formate dehydrogenase activities were detected in cell-free extracts using benzyl viologen as an artificial electron acceptor [17]. On the other hand, genes annotated as pyruvate:formate lyase or formate C-acetyltransferase were identified in the genome (Ga0116910_1004109, Ga0116910_100860, Ga0116910_102934 and Ga0116910_102935), but no activity for a possible pyruvate:formate lyase could be detected [Patil et al., unpublished results]. This indicates that the genomic information is sometimes insufficient to predict metabolic pathways. Thus, further biochemical and proteomics studies would be needed to investigate and confirm the gluconate and glucose fermentation pathway utilized by this bacterium.

Microcompartments and fucose utilization

The genome of A. acetethylicum strain GluBS11T harbors five genes that putatively code for bacterial microcompartment shell proteins. Four of these genes are annotated as “BMC-domain-containing protein” (Ga0116910_1005148, Ga0116910_1005149, Ga0116910_1005150 and Ga0116910_1005151), and one gene is annotated as “Carboxysome shell and ethanolamine utilization microcompartment protein CcmL/EutN” (Ga0116910_1005155). Microcompartments are protein complexes that form discrete spaces within the cell, thus enabling enzyme reactions that either produce toxic intermediates or require accumulation of a certain metabolite, e.g., the ethanolamine utilization microcompartment in Salmonella typhimurium ATCC 13311 T or the carboxysomes in cyanobacteria [50, 51]. An IMG gene search for microcompartments and subsequent comparison to other genomes using the IMG Gene Ortholog Neighborhoods viewer, revealed that the microcompartment genes in A. acetethylicum strain GluBS11T are located in a putative operon that also contains genes associated with fucose utilization in Clostridium phytofermentans ATCC 700394 T [52]. Fucose, a deoxyhexose derived from plant biomass degradation, can be fermented to propionate, propanol, mixed acids, and ethanol by C. phytofermentans ATCC 700394 T , and the responsible genes are located in two different operons in this organism [52]. Initially, fucose is converted to fuculose-phosphate by fucose mutarotase, fucose isomerase and fucose kinase (Cphy_3153 – Cphy_3155); [52]. Likewise, the orthologs in A. acetethylicum strain GluBS11T are located in a similar operon (L-fucose isomerase Ga0116910_100812, rhamnulokinase/L-fuculokinase Ga0116910_100813 and L-fucose mutarotase Ga0116910_100815). Fuculose-phosphate is then further degraded to lactaldehyde and dihydroxyacetone-phosphate by fuculose-phosphate aldolase (Ga0116910_102223 in A. acetethylicum strain GluBS11T, Cphy_1177 in C. phytofermentans ). Dihydroxyacetone phosphate can then be processed through glycolysis, while lactaldehyde is reduced to 1,2-propanediol with NADH. 1,2-propanediol is then disproportionated in microcompartments to propionate and propanol by 1,2-propanediol oxidoreductase (Cphy_1185, Ga0116910_1005154), 1,2-propanediol dehydratase (Cphy_1174, Ga0116910_100557 - Ga0116910_100559 in a different area of the genome), propionaldehyde dehydrogenase (Cphy_1178, Ga0116910_1005146), phosphate propanoyl transferase (Cphy_1183, Ga0116910_1005152), acetate/propionate kinase (Cphy_1327, Ga0116910_104214, Ga0116910_1001586, or Ga0116910_103636) and propanol dehydrogenase (Cphy_1179, Ga0116910_1005147). Rhamnose can be degraded in a similar way by C. phytofermentans ATCC 700394 T , and the respective genes leading to lactaldehyde and dihydroxyacetone-phosphate were also identified in the genome of A. acetethylicum strain GluBS11T (L-rhamnose mutarotase Ga0116910_10513, L-rhamnose isomerase Ga0116910_1001301, rhamnulokinase/L-fuculokinase Ga0116910_100813) [52]. However, earlier results demonstrated that rhamnose cannot be utilized by A. acetehylicum strain GluBS11T [52]. Even though the genes for fucose degradation are present in the genome, it is still doubtful whether this sugar can serve as a growth-supporting substrate for strain GluBS11T.

Conclusions

Taken together, the draft genome sequence of A. acetethylicum strain GluBS11T expands our view on the metabolic capacities of this sugars and glycerol-fermenting bacterium. The genome sequence gives us insights into the putative enzymes involved in the pathway of glucose and gluconate (including glycerol) fermentation, and provides a brief summary of the key reactions involved. Lastly, the hypotheses concerning the glucose and gluconate fermentation pathways based on genomic data are still preliminary, and additional biochemical and functional proteomic studies will be necessary for pathway confirmation and further insights.

Abbreviations

CDS:

Coding DNA sequence

COG:

Clusters of orthologous groups

CTAB:

Cetyl trimethyl ammonium bromide

KEGG:

Kyoto encyclopedia of genes and genomes

MEGA:

Molecular evolutionary genetics analysis

NADH:

Nicotinamide adenine dinucleotide reduced

References

  1. Patil Y, Junghare M, Pester M, Müller N, Schink B. Characterization and phylogeny of Anaerobium acetethylicum gen. nov., sp. nov., a strictly anaerobic gluconate-fermenting bacterium isolated from a methanogenic bioreactor. Int J Syst Evol Microbiol. 2015;65:3289–96.

    Article  CAS  PubMed  Google Scholar 

  2. Rainey FA. Family V. Lachnospiraceae fam., nov. In: De Vos P, Garrity GM, Jones D, Krieg NR, Ludwig W, Rainey FA, Schleifer KH, Whitman WB, editors. Bergey’s Manual of Systematic Bacteriology, vol. volume 3. 2nd ed. New York: Springer; 2009. p. 921.

    Google Scholar 

  3. Rainey FA. Class II Clostridia class., nov. In: De Vos P, Garrity GM, Jones D, Krieg NR, Ludwig W, Rainey FA, Schleifer KH, Whitman WB, editors. Bergey’s Manual of Systematic Bacteriology, vol. volume 3. 2nd ed. New York: Springer; 2009. p. 736–1297.

    Google Scholar 

  4. Prévot AR. In: Hauderoy P, Ehringer G, Guillot G, Magrou J, Prévot AR, Rosset D, Urbain A, editors. Dictionnaire des Bactéries Pathogènes. 2nd ed. Paris: Masson et Cie; 1953. p. 1–692.

    Google Scholar 

  5. Collins MD, Lawson PA, Willems A, Cordoba JJ, Fernandez-Garayzabal J, Garcia P, Cai J, Hippe H, Farrow JAE. The phylogeny of the genus Clostridium: proposal of five new genera and eleven new species combinations. Int J Syst Bacteriol. 1994;44:812–26.

    Article  CAS  PubMed  Google Scholar 

  6. Stackebrandt E, Kramer I, Swiderski J, Hippe H. Phylogenetic basis for a taxonomic dissection of the genus Clostridium. FEMS Immunol Med Microbiol. 1999;24:253–8.

    Article  CAS  PubMed  Google Scholar 

  7. Kittelmann S, Seedorf H, Walters WA, Clemente JC, Knight R, Gordon JI, Janssen PH. Simultaneous amplicon sequencing to explore co-occurrence patterns of bacterial, archaeal and eukaryotic microorganisms in rumen microbial communities. PLoS One. 2013;8:47879.

    Article  Google Scholar 

  8. Gosalbes MJ, et al. Metatranscriptomic approach to analyze the functional human gut microbiota. PLoS One. 2011;6:17447.

    Article  Google Scholar 

  9. Cotta M, Forster R. The Family Lachnospiraceae, Including the Genera Butyrivibrio, Lachnospira and Roseburia. In: Dworkin M, Falkow S, Rosenberg E, Schleifer KH, Stackebrandt E, editors. The Prokaryotes. Bacteria: Firmicutes, Cyanobacteria a handbook on the biology of bacteria. New York: Springer; 2006. p. 1002–21.

    Chapter  Google Scholar 

  10. Biddle A, Stewart L, Blanchard J, Leschine S. Untangling the genetic basis of fibrolytic specialization by Lachnospiraceae and Ruminococcaceae in diverse gut communities. Diversity. 2013;5:627–40.

    Article  Google Scholar 

  11. Barcenilla A, Pryde SE, Martin JC, Duncan SH, Stewart CS, Henderson C, Flint HJ. Phylogenetic relationships of butyrate-producing bacteria from the human gut. Appl Environ Microbiol. 2000;66:1654–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Duncan SH, Barcenilla A, Stewart CS, Pryde SE, Flint HJ. Acetate utilization and butyryl coenzyme A (CoA): acetate-CoA transferase in butyrate-producing bacteria from the human large intestine. Appl Environ Microbiol. 2002;68:5186–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Varel VH, Tanner RS, Woese CR. Clostridium herbivorans sp. nov., a cellulolytic anaerobe from the pig intestine. Int J Syst Bacteriol. 1995;45:490–4.

    Article  CAS  PubMed  Google Scholar 

  14. Sleat R, Mah RA. Clostridium populeti sp. nov., a cellulolytic species from a woody-biomass digestor. Int J Syst Bacteriol. 1985;35:160–3.

    Article  Google Scholar 

  15. van Gylswyk NO. Fusobacterium polysaccharolyticum sp. nov., a Gram-negative rod from the rumen that produces butyrate and ferments cellulose and starch. J Gen Microbiol. 1980;116:157–63.

    PubMed  Google Scholar 

  16. van Gylswyk NO, Morris EJ, Els HJ. Sporulation and cell wall structure of Clostridium polysaccharolyticum comb. nov. (formerly Fusobacterium polysaccharolyticum). J Gen Microbiol. 1980;121:491–3.

    Google Scholar 

  17. Patil Y, Junghare M, Müller N. Fermentation of glycerol by Anaerobium acetethylicum and its potential use in biofuel production. Microbiol Biotechnol. 2017;10:203–17.

  18. Homann T, Tag C, Biebl H, Deckwer WD, Schink B. Fermentation of glycerol to 1,3-propanediol by Klebsiella and Citrobacter strains. Appl Microbiol Biotechnol. 1990;33:121–6.

    Article  CAS  Google Scholar 

  19. Gibbons NE, Murray RGE. Proposals concerning the higher taxa of bacteria. Int J Syst Bacteriol. 1978;28:1–6.

    Article  Google Scholar 

  20. Kumar S, Stecher G, Tamura K. MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol Biol Evol. 2016;33:1870–4.

    Article  CAS  PubMed  Google Scholar 

  21. Rzhetsky A, Nei M. A simple method for estimating and testing minimum-evolution trees. Mol Biol Evol. 1992;9:945–67.

    CAS  Google Scholar 

  22. Jukes TH, Cantor CR. Evolution of protein molecules. In: Munro HN, editor. Mammalian protein metabolism. Academic Press: New York; 1969. p. 21–132.

    Chapter  Google Scholar 

  23. Saitou N, Nei M. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol. 1987;4:406–25.

    CAS  PubMed  Google Scholar 

  24. Kyrpides NC, Hugenholtz P, Eisen JA, Woyke T, Göker M, Parker CT, et al. Genomic Encyclopedia of Bacteria and Archaea: sequencing a myriad of type strains. PLoS Biol. 2014;12:1001920.

    Article  Google Scholar 

  25. Whitman WB, Woyke T, Klenk H-P, et al. Genomic Encyclopedia of Bacterial and Archaeal type strains, Phase III: the genomes of soil and plant-associated and newly described type strains. Stand Genomic Sci. 2015;10:26.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Reddy TBK, Thomas AD, Stamatis D, Bertsch J, Isbandi M, Jansson J, Mallajosyula J, Pagani I, Lobos EA, Kyrpides NC. The Genomes OnLine Database (GOLD) v. 5: a metadata management system based on a four level (meta)genome project classification. Nucleic Acids Res. 2015;43:1099–106.

    Article  Google Scholar 

  27. Markowitz VM, Chen IM, Palaniappan K, Chu K, Szeto E, Pillay M, Ratner A, Huang J, Woyke T, Huntemann M, et al. IMG 4 version of the integrated microbial genomes comparative analysis system. Nucleic Acids Res. 2014;42:560–7.

    Article  Google Scholar 

  28. Mavromatis K, Land ML, Brettin TS, Quest DJ, Copeland A, Clum A, Goodwin L, Woyke T, Lapidus A, Klenk HP, et al. The fast changing landscape of sequencing technologies and their impact on microbial genome assemblies and annotation. PLoS One. 2012;7:48837.

    Article  Google Scholar 

  29. Porebski S, Bailey L, Baum B. Modification of a CTAB DNA extraction protocol for plants containing high polysaccharide and polyphenol components. Plant Mol Biol Rep. 1997;15:8–15.

    Article  CAS  Google Scholar 

  30. Junghare M, Patil Y, Schink B. Draft genome sequence of a nitrate-reducing, o-phthalate degrading bacterium, Azoarcus sp. strain PA01T. Stand Genomic Sci. 2015;10:90.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Bennett S. Solexa Ltd. Pharmacogenomics. 2004;5:433–8.

    Article  PubMed  Google Scholar 

  32. Bushnell B. BBTools software package. https://sourceforge.net/projects/bbmap/

  33. Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, Lesin VM, Nikolenko SI, Pham S, Prjibelski AD, Pyshkin AV, Sirotkin AV, Vyahhi N, Tesler G, Alekseyev MA, Pevzner PA. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J Comput Biol. 2012;19:455–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Hyatt D, Chen GL, Locascio PF, Land ML, Lar-imer FW, Hauser LJ. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinformatics. 2010;11:119.

    Article  PubMed  PubMed Central  Google Scholar 

  35. Huntemann M, Ivanova NN, Mavromatis K, Tripp HJ, Paez-Espino D, Palaniappan K, Szeto E, Pillay M, Chen IM, Pati A, et al. The standard operating procedure of the DOE-JGI Microbial Genome Annotation Pipeline (MGAP v.4). Stand Genomic Sci. 2015;10:86.

    Article  PubMed  PubMed Central  Google Scholar 

  36. Lowe TM, Eddy SR. tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res. 1997;25:955–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Pruesse E, Quast C, Knittel K, Fuchs BM, Ludwig W, Peplies J, et al. SILVA: a comprehensive online resource for quality checked and aligned ribosomal RNA sequence data compatible with ARB. Nucleic Acids Res. 2007;35:7188–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Nawrocki EP, Eddy SR. Infernal 1.1: 100-fold faster RNA homology searches. Bioinformatics. 2013;29:2933–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. 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;25:2271–8.

    Article  CAS  PubMed  Google Scholar 

  40. Grissa I, Vergnaud G, Pourcel C. The CRISPR db database and tools to display CRISPRs and to generate dictionaries of spacers and repeats. BMC Bioinformatics. 2007;8:172.

    Article  PubMed  PubMed Central  Google Scholar 

  41. González JM, Fernández-Gómez B, Fernàndez-Guerra A, Gómez-Consarnau L, Sánchez O, Coll-Lladó M, Del Campo J, Escudero L, Rodríguez-Martínez R, Alonso-Sáez L, Latasa M, Paulsen I, Nedashkovskaya O, Lekunberri I, Pinhassi J, Pedrós-Alió C. Genome analysis of the proteorhodopsin-containing marine bacterium Polaribacter sp. MED152 (Flavobacteria). Proc Natl Acad Sci USA. 2008;105:8724–9.

    Article  PubMed  PubMed Central  Google Scholar 

  42. Grupe H, Gottschalk G. Physiological events in Clostridium acetobutylicum during the shift from acidogenesis to solventogenesis in continuous culture and presentation of a model for shift induction. Appl Environ Microbiol. 1992;58:3896–902.

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Al-Awqati Q. Proton-translocating ATPases. Annu Rev Cell Biol. 1986;2:179–99.

    Article  CAS  PubMed  Google Scholar 

  44. Ramachandran S, Fontanille P, Pandey A, Larroche C. Gluconic acid. properties, applications and microbial production. Food Technol Biotechnol. 2006;44:185–95.

    CAS  Google Scholar 

  45. Andreesen JR, Gottschalk G. The occurrence of a modified Entner-Doudoroff pathway in Clostridium aceticum. Arch Mikrobiol. 1969;69:160–70.

    Article  CAS  PubMed  Google Scholar 

  46. Eisenberg RC, Dobrogosz WJ. Gluconate metabolism in Escherichia coli. J Bacteriol. 1967;93:941–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Servinsky MD, Liu S, Gerlach ES, Germane KL, Sund CJ. Fermentation of oxidized hexose derivatives by Clostridium acetobutylicum. Microb Cell Fact. 2014;13:139.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. Basic local alignment search tool. J Mol Biol. 1990;215:403–10.

    Article  CAS  PubMed  Google Scholar 

  49. Bender R, Andreesen JR, Gottschalk G. 2-Keto-3-deoxygluconate, an intermediate in the fermentation of gluconate by Clostridia. J Bacteriol. 1971;107:570–3.

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Garsin DA. Ethanolamine utilization in bacterial pathogens: roles and regulation. Nat Rev Microbiol. 2010;8:290–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Rae BD, Long BM, Whitehead LF, Förster B, Badger MR, Price GD. Cyanobacterial carboxysomes: microcompartments that facilitate CO2 fixation. J Mol Microbiol Biotechnol. 2013;23:300–7.

    Article  CAS  PubMed  Google Scholar 

  52. Petit E, LaTouf WG, Coppi MV, Warnick TA, Currie D, Romashko I, Deshpande S, Haas K, Alvelo-Maurosa JG, Wardman C, Schnell DJ, Leschine SB, Blanchard JL. Involvement of a bacterial microcompartment in the metabolism of fucose and rhamnose by Clostridium phytofermentans. PLoS One. 2013;8:54337.

    Article  Google Scholar 

  53. Field D, Garrity GM, Gray T, Morrison N, Selengut J, Sterk P, et al. The minimum information about a genome sequence (MIGS) specification. Nat Biotechnol. 2008;26:541–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. 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;87:4576–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Murray RGE. The higher taxa, or, a place for everything…? In: Krieg NR, Hol JG, editors. Bergey's Manual of Systematic Bacteriology, vol. 1. Baltimore: Williams & Wilkins Co; 1984. p. 31–4.

    Google Scholar 

  56. Euzéby J. List of new names and new combinations previously effectively, but not validly, published. List no. 132. Int J Syst Evol Microbiol. 2010;60:469–72.

    Article  Google Scholar 

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

    Article  Google Scholar 

  58. Ashburner M, Ball CA, Blake JA, Botstein D, Butler H, Cherry JM, Davis AP, Dolinski K, Dwight SS, Eppig JT, Harris MA, Hill DP, Issel-Tarver L, Kasarskis A, Lewis S, Matese JC, Richardson JE, Ringwald M, Rubin GM, Sherlock G. Gene ontology: tool for the unification of biology. The gene ontology consortium. Nat Gene. 2000;25:25–9.

    Article  CAS  Google Scholar 

  59. Nei M, Kumar S. Molecular Evolution and Phylogenetics. New York: Oxford University Press; 2000.

    Google Scholar 

  60. Felsenstein J. Confidence limits on phylogenies: an approach using the bootstrap. Evol. 1985;39:783–91.

    Article  Google Scholar 

Download references

Acknowledgements

YP thanks the LGFG scholarship funding program of the University of Konstanz for providing scholarship during this research work. The authors appreciate the service of the Electron Microscopy Center of the University of Konstanz.

Funding

During this research YP was funded by a LGFG PhD scholarship. The genome sequencing was performed under the auspices of the US Department of Energy’s Office of Science, Biological and Environmental Research Program, and by the University of California, Lawrence Berkeley National Laboratory under contract No. DE-AC02-05CH11231.

Authors’ contributions

YP, MJ, NM and BS initiated the project and YP performed DNA preparation. MJ and YP performed the comparative genomics, investigated the genome for general metabolic features and fermentation pathways. MJ, YP and NM drafted the manuscript that was critically reviewed and corrected by BS, NM, WW, NS and NK, respectively. MH, AC, MP, KP, NV, NM, DS, TBKR, CD, NI, and TW performed the technical work for sequencing, assembly and annotation of the genome. All authors read and approved the final manuscript.

Competing interests

The authors declare that they have no competing interests.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Nicolai Müller, Bernhard Schink or Madan Junghare.

Additional files

Additional file 1: Figure S1.

Gel electrophoresis of genomic DNA isolated from GluBS11T cells grown with gluconate. (TIF 159 kb)

Additional file 2: Table S1.

IMG annotated functions of selected putative key enzymes involved in the metabolic pathways identified in the draft genome sequence of A. acetethylicum strain GluBS11T. (DOCX 15 kb)

Additional file 3: Table S2.

Putative transporters identified in the draft genome of A. acetethylicum GluBS11T. (DOCX 14 kb)

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Patil, Y., Müller, N., Schink, B. et al. High-quality-draft genome sequence of the fermenting bacterium Anaerobium acetethylicum type strain GluBS11T (DSM 29698). Stand in Genomic Sci 12, 24 (2017). https://doi.org/10.1186/s40793-017-0236-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40793-017-0236-4

Keywords