Skip to main content
  • Research article
  • Open access
  • Published:

Evolution from a respiratory ancestor to fill syntrophic and fermentative niches: comparative genomics of six Geobacteraceae species

Abstract

Background

The anaerobic degradation of organic matter in natural environments, and the biotechnical use of anaerobes in energy production and remediation of subsurface environments, both require the cooperative activity of a diversity of microorganisms in different metabolic niches. The Geobacteraceae family contains members with three important anaerobic metabolisms: fermentation, syntrophic degradation of fermentation intermediates, and anaerobic respiration.

Results

In order to learn more about the evolution of anaerobic microbial communities, the genome sequences of six Geobacteraceae species were analyzed. The results indicate that the last common Geobacteraceae ancestor contained sufficient genes for anaerobic respiration, completely oxidizing organic compounds with the reduction of external electron acceptors, features that are still retained in modern Geobacter and Desulfuromonas species. Evolution of specialization for fermentative growth arose twice, via distinct lateral gene transfer events, in Pelobacter carbinolicus and Pelobacter propionicus. Furthermore, P. carbinolicus gained hydrogenase genes and genes for ferredoxin reduction that appear to permit syntrophic growth via hydrogen production. The gain of new physiological capabilities in the Pelobacter species were accompanied by the loss of several key genes necessary for the complete oxidation of organic compounds and the genes for the c-type cytochromes required for extracellular electron transfer.

Conclusion

The results suggest that Pelobacter species evolved parallel strategies to enhance their ability to compete in environments in which electron acceptors for anaerobic respiration were limiting. More generally, these results demonstrate how relatively few gene changes can dramatically transform metabolic capabilities and expand the range of environments in which microorganisms can compete.

Background

The global carbon cycle and the production of a number of biofuels depends on the cooperative interaction of a physiological diversity of anaerobic microorganisms. These include microorganisms that ferment complex substrates to simpler molecules; respiratory microorganisms that convert fermentation products to carbon dioxide and/or methane; and syntrophic organisms that anaerobically oxidize some fermentation products and other substrates in reactions that are only thermodynamically feasible when respiratory organisms consume the syntrophs' products. The family Geobacteraceae has representatives of all three metabolic groups, providing an opportunity to evaluate how microorganisms might evolve to fill these various niches in anaerobic ecosystems.

Within the Geobacteraceae, Geobacter and Desulfuromonas are specialists in the complete oxidation of organic compounds to carbon dioxide coupled to the reduction of insoluble, extracellular electron acceptors [1]. In aquatic sediments and submerged soils they influence the cycling of carbon and metals by oxidizing acetate, the central fermentation intermediate in anaerobic environments, with the reduction of iron and manganese oxides [1, 2]. In hydrocarbon-contaminated subsurface environments Geobacter species can play an important role in the bioremediation of aromatic hydrocarbons by oxidizing the contaminants with the reduction of Fe(III) oxides naturally present in the subsurface [3, 4]. This is a process which can be artificially stimulated [5, 6]. Another bioremediation application involving Geobacter species is to add acetate to uranium-contaminated subsurface environments. This stimulates the growth of Geobacter species which reduce soluble, mobile U(VI) to insoluble U(IV) [7] and thus immobilize the uranium in situ [4, 8]. The ability of Geobacter and Desulfuromonas species to oxidize organic compounds with electron transfer to graphite electrodes provides a convenient method for harvesting electricity from aquatic sediments and organic wastes to power electronic devices [9–11].

In contrast, the primary niche of Pelobacter species is methanogenic environments, functioning either as fermentative microorganisms or living in syntrophic associations with methanogens by partially oxidizing organic compounds to hydrogen and acetate which the methanogens consume [12, 13]. Pelobacter species cannot oxidize acetate or other organic compounds completely to carbon dioxide and are ineffective in electron transfer to metals [14] or electrodes [15]. Therefore, not only the natural environments, but also the technological applications of Geobacter/Desulfuromonas versus Pelobacter species are quite different.

Here we report on the sequenced genomes of six members of the Geobacteraceae family: Geobacter sulfurreducens [16], Geobacter metallireducens, Geobacter uraniireducens, Pelobacter propionicus, Pelobacter carbinolicus, and Desulfuromonas acetoxidans. We use comparative genomics in an effort to better understand the enzymes involved in these unusual metabolisms, and to provide insight in to the selective pressures that may have contributed to their evolution.

Results and discussion

Identification of protein families in the six genomes

The general features of each of the six genomes are presented in Table 1. Orthologous proteins, those predicted to have similar functions in the different organisms, were identified by Markov clustering of sets of reciprocal best BLAST matches [17]. Using all 21,716 protein coding genes in the six genomes (see Additional file 1), 3,696 clusters with members from at least two genomes were defined (see Additional file 2). 15,207 proteins (70.0%) were classified into clusters. A functional role was predicted for each cluster using the G. sulfurreducens in silico model annotation [18] and COG categorization [19].

Table 1 General features of the Geobacteraceae genomes

The pattern of conservation of the proteins families across all species (the phyletic pattern) was determined (see Additional files 2 and 3). By far the most common pattern was conservation across all species, 5,345 proteins were members of clusters that included at least one protein from each genome (see Additional file 3). The second most common pattern was conservation in the Geobacter species and P. propionicus only (1,158 proteins). Proteins unique to the respiratory or fermentative species were relatively rare. Only 370 proteins were found in the respiratory species but not the fermentative species (see Additional file 4), and only 260 proteins were found in the fermentative species but not the respiratory species (see Additional file 5).

Whole-genome phylogeny

A phylogeny of the family was constructed using the 481 protein families that had a single protein from each of the six genomes and the outgroup species Anaeromyxobacter dehalogenans (see Additional file 6). These proteins had diverse functions, including information storage, metabolism, cell signaling, and unknown. The proteins from each genome were concatenated then aligned, and this alignment was used to create a Bayseian model of the phylogeny (Figure 1). This analysis confirmed that, as housekeeping-gene phylogeny suggests [20], the Geobacteraceae can be divided into two clades, and that the two fermentative Pelobacter species are not most closely related to each other, instead there is one Pelobacter species in each clade of the respiratory species (Figure 1).

Figure 1
figure 1

Genome-based Geobacteraceae phylogeny. Bayesian inference of the phylogenetic tree of the six Geobacteraceae species discussed in the text, using another Deltaproteobacterial species, Anaeromyxobacter dehalogenans, as the outgroup. The tree was based on a concatenation of the proteins in the 481 families that had exactly one ortholog conserved in each of the seven genomes (see Additional file 6). Values at branch points are posterior probabilities.

Thus, based on phylogeny, there are at least three possibilities for how modern Geobacteraceae evolved to fill multiple metabolic niches. The ancestor may have been adapted to a fermentative/syntrophic lifestyle with no reliance on external electron acceptors. In this scenario, the Geobacter and Desulfuromonas species would have independently evolved strategies for the complete oxidation of organic compounds coupled to Fe(III) reduction. Alternatively, the ancestor may have been a respiratory microorganism, and the Pelobacter species independently evolved strategies for fermentation and syntrophy and lost the ability to transfer electrons out of the cell. Another option was that the ancestor may have had both types of metabolism, and the various branches of the family lost certain abilities. We investigated these possibilities by analyzing the conservation and evolutionary history of genes for energy metabolism.

Conservation of pathways for acetate oxidation

In the respiratory species G. sulfurreducens, G. metallireducens, G. uraniireducens, and D. acetoxidans, acetate is the primary electron donor and it is oxidized via the TCA cycle, generating NADH, NADPH, and reduced ferredoxin (Figure 2) [21–23]. In contrast, Pelobacter species are incapable of acetate oxidation [12]. One protein family predicted to be acetate transporters (GSU0518) [24] was conserved in all of the acetate-oxidizing species (see Additional file 7). However, neither this protein nor any of the other acetate transporters found in G. sulfurreducens (GSU1068, GSU1070, and GSU2352) were conserved in either of the Pelobacter species (see Additional file 7). Twenty one genes encode the enzymes of the TCA cycle used by G. sulfurreducens [18] (Figure 2, see Additional file 7). There was full conservation in all of the respiratory species of at least one copy of each of these enzymes (see Additional file 7). The one exception was the malate dehydrogenase in D. acetoxidans. There was no gene predicted to encode a malate dehydrogenase in the incomplete version of the genome, but this enzyme activity is found in D. acetoxidans [23]. As with the acetate transporters, not all of the genetic redundancy seen in G. sulfurreducens in the enzymes of the TCA cycle was conserved across the family. Only one copy of the aconitase (GSU1660) and the keto/oxoacid ferredoxin oxidoreductase (GSU1859–GSU1862) were conserved in all the respiratory species (see Additional file 7).

Figure 2
figure 2

The pathways of energy metabolism conserved in the respiratory Geobacteraceae species. Shown are the pathways, based those characterized in G. sulfurreducens, for acetate activation and oxidation via the TCA cycle in the cytoplasm, for inner membrane oxidation of TCA cycle products coupled with electron and proton transport, and for ATP generation. The genes encoding the enzymes of these pathways and their conservation pattern in all of the Geobacteraceae genomes are given in see Additional file 7. The enzymes are colored black if there were orthologs for every subunit in all of the respiratory species. The asterisk indicates that malate dehydrogenase activity has been shown in D. acetoxidans, though genes for the enzyme were not found in the draft sequence of the genome. The pathway for electron transfer through the periplasm and out to the external electron acceptor is not well characterized in these species, and is represented here by the three cytochromes known to be required in vivo in G. sulfurreducens that are also conserved in all of the respiratory species.

The Pelobacter species contained orthologs to many of the TCA cycle enzymes (see Additional file 7), but there were two notable exceptions. NADPH metabolism appears to catalyzed by different enzymes in the Pelobacter species. A monomeric-type isocitrate dehydrogenase (GSU1465) was found in all of the respiratory species (see Additional file 8), but was not conserved in either Pelobacter species (see Additional file 7). In acetate oxidizers, this reaction is the primary source of NADPH in the cell [25]. Instead, the Pelobacter species contained non-orthologous isocitrate dehydrogenase genes (Ppro_0452 and Pcar_1038 in cluster 3107) of the more common, homodimeric type [25]. In addition, both Pelobacter species lacked the dehydrogenase (GSU0509–GSU0510) believed to transfer the electrons from NADPH into the electron transport chain [26]. This enzyme was conserved in all four of the respiratory species (see Additional file 7). These changes suggest that the NADPH produced by the Pelobacter species may not be used primarily as a source of electrons for energy metabolism, as it is for the respiratory species. Furthermore, P. carbinolicus contained a non-orthologous fumarase enzyme (Pcar_0324) predicted to be of the class II, Fe(III)-free type [27], rather than the class I, Fe(III) type found in the other Geobacteraceae (GSU0994).

Conservation of pathways for extracellular electron transfer

In the respiratory Geobacteraceae species, the reducing equivalents from the TCA cycle are passed to inner membrane quinones via NADH dehydrogenase (Figure 2) [21–23]. This is predicted to be the only step in the electron transport chain where protons are pumped across the inner membrane for ATP generation [18]. G. sulfurreducens encodes two NADH dehydrogenase operons, one with 12 subunits and one with 14 (see Additional file 7). The 14-subunit enzyme is conserved in all of the respiratory species and P. propionicus (see Additional file 7). In addition, P. propionicus appears to have recently duplicated this enzyme, there are three virtually identical copies in the genome (Ppro_0628–Ppro_0641, Ppro_1623–Ppro_1636, Ppro_3180–Ppro_3193). In contrast, P. carbinolicus lacks the 14-subunit enzyme, but encodes an ortholog to the 12-subunit enzyme (see Additional file 7). This conservation pattern indicates that the 14-subunit enzyme may be the more important for inner membrane proton and electron transport; it is conserved in the four respiratory species, and in the only fermentative species predicted to use an NADH dehydrogenase (as described below and in Figure 3).

Figure 3
figure 3

The pathways of energy metabolism in Pelobacter propionicus and the differences from respiratory Geobacteraceae species. Shown are the pathways of butanediol and acetoin fermentation used by P. propionicus. Vertically-inherited enzymes with orthologs in the other Geobacteraceae are shown in black. Both the vertically-inherited NADH dehydrogenase and ATP synthase have been duplicated in P. propionicus and these are shown in dark grey. Enzymes acquired by lateral gene transfer are shown in red: 1) butanediol dehydrogenase, 2) acetoin dehydrogenase, 3) pyruvate:ferredoxin oxidoreductase, 4) transcarboxylase, and 5) methylmalonyl-CoA mutase, and an ATP synthase. Enzymes conserved in the respiratory Geobacteraceae but lost in P. propionicus are shown in white. Reactions with multiple enzymes encoded in the genome for which there are diverse conservation patterns are shown as dashed arrows.

The final part of the electron transport chain in the respiratory species is transfer of electrons from the inner membrane to cytochromes for transport to the cell surface. The differences between the respiratory and fermentative species are substantial in this part of the pathway. The genomes of all four of the respiratory species contained at least one copy of a putative cytochrome bc complex (GSU2932–GSU2934), which is predicted to transfer electrons from membrane-bound quinones to periplasmic cytochromes [28] (Figure 2, see Additional file 7). Neither of the Pelobacter species had any orthologs to this complex. C-type cytochromes span the periplasm and outer membrane in the respiratory species [1]. All of the Geobacter and Desulfuromonas genomes contained more than 100 ORFs that have the heme-binding motif [29] characteristic of c-type cytochromes (see Additional file 1). Both Pelobacter genomes had far fewer cytochrome genes, ca. 40 per genome (see Additional file 1). Many of the proteins found only in the respiratory species were multi-heme, c-type cytochromes (see Additional file 4), whereas none of the proteins specific to the fermentative species were multi-heme (see Additional file 5). In addition, both Pelobacter species lacked most of the specific cytochromes required for electron transfer in G. sulfurreducens in vivo. These include those conserved in all of the respiratory species: MacA, a c-type cytochrome associated with the inner membrane [30] and OmcB, an outer-membrane cytochrome [31].

Lateral transfer of fermentation genes

Thus, while the Pelobacter species contained many vertically-inherited genes for respiratory metabolism, several key enzymes are different or missing in both species. To investigate evolution of the fermentative and syntrophic metabolism, Pelobacter-specific genes were analyzed, and genes originating from lateral gene transfer were identified using a combination of phylogenetic and BLAST-based analysis (see Additional file 9). Both Pelobacter species catabolize butanediol and acetoin [12]. The butanediol dehydrogenase (Bdh, Pcar_0330) and acetoin dehydrogenase (AcoABCL, Pcar_0343–Pcar0346), which catalyze the initial steps in the metabolism of these compounds, have been characterized in P. carbinolicus [32]. P. propionicus has genes with high sequence similarity to these, but the operon structure of the putative acetoin dehydrogenase included a duplication of the A and B subunits (Bdh, Ppro_1043 and AcoABCABL, Ppro_1024–Ppro1029, see Additional file 1). Phylogenetic analysis showed that the most closely related enzymes are not from Geobacter or Desulfuromonas, nor any other delta-Proteobacteria species (see Additional file 10). Instead the Pelobacter genes were most closely related to the butanediol catabolic genes from Pseudomonas [33] and Gram-positive species, indicating that these genes originated from lateral gene transfer (see Additional file 10).

The subsequent steps in the catabolism of the acetyl-CoA and acetaldehyde from acetoin are markedly different in the two Pelobacter species. The pathway in P. propionicus is cyclic and requires membrane-bound electron transport enzymes (Figure 3) [34]. Several of the key enzymes were predicted to have been acquired by lateral gene transfer. Pyruvate:ferredoxin oxidoreductase is required to convert acetyl-CoA to pyruvate (Figure 3), and there were two heterotetrameric pyruvate:ferredoxin oxidoreductases from lateral gene transfer. One (Ppro_0322–Ppro_0325, see Additional file 1) is most closely related to the enzymes from Clostridia tetani and Thermotoga species, and the other (Ppro_0469–Ppro_0472, see Additional file 1) is most similar to the enzyme in Syntrophobacter and several archaeal species.

Pyruvate is then converted to succinyl-CoA to regenerate NAD+ without the need for a cooperating hydrogen-oxidizing species (Figure 3). P. propionicus has orthologs of four TCA cycle enzymes and the NADH dehydrogenase from the respiratory Geobacteraceae that could carry out these reactions (Figure 3, see Additional file 7). Then, propionate is generated via a methylmalonyl-CoA mutase and transcarboxylase (Figure 3). Both of these enzyme were predicted to have been acquired from lateral gene transfer. The mutase (Ppro_1284–Ppro_1285, see Additional file 1) is related to enzymes from Chlorflexus and Geobacillus species, and the transcarboxylase (Ppro_0033–Ppro_0034, see Additional file 1) is related to that crystallized from Propionibacterium freudenreichii [35].

Thus, it appears that P. propionicus uses a mosaic of vertically inherited and laterally acquired genes for fermentation. Interestingly, the pathway uses a complete, vertically inherited, respiratory electron transport chain – the succinate dehydrogenase operates in reverse as a fumarate reductase, accepting reducing equivalents from the NADH dehydrogenase [34]. This is equivalent to that used by Geobacter species growing by fumarate respiration [18, 36].

Differences between Pelobacter species associated with syntrophy

P. carbinolicus ferments via much simpler pathway of cytoplasmic enzymes (Figure 4). As described above, the butanediol dehydrogenase and acetoin dehydrogenase appear to have been acquired by lateral gene transfer. NAD+ is regenerated either by ethanol production or by proton reduction to hydrogen, which requires a syntrophic partner and ATP is made by substrate level phosphorylation by an acetate kinase [12].

Figure 4
figure 4

The pathways of energy metabolism in Pelobacter carbinolicus and the differences from respiratory Geobacteraceae species. Shown are the pathways of butanediol and acetoin ethanol fermentation and syntrophy used by P. carbinolicus. Vertically-inherited enzymes with orthologs in the other Geobacteraceae are shown in black. The vertically-inherited ATP synthase has been duplicated in P. carbinolicus and is shown in dark grey. Enzymes acquired by lateral gene transfer are shown in red: 1) butanediol dehydrogenase, 2) acetoin dehydrogenase, 3) hydrogenases, the RNF electron transfer complex, and possibly an ATP synthase (see text). Enzymes conserved in the respiratory Geobacteraceae but lost in P. carbinolicus are shown in white. Reactions with multiple enzymes encoded in the genome for which there are diverse conservation patterns are shown as dashed arrows.

Some Pelobacter species live in syntrophic associations with methanogens by partially oxidizing organic compounds to hydrogen and acetate which the methanogens consume [12, 13]. Both Pelobacter species lacked orthologs to any of the Geobacter-type [37] hydrogenases (GSU0121–GSU0123, GSU0782–GSU0785, GSU2417–GSU2420 or GSU2718–GSU2722, see Additional file 1). Furthermore, the hydrogenases in the two Pelobacter species are dissimilar. This may reflect the fact that P. carbinolicus is capable of growing syntrophically with a hydrogen-consuming partner, most typically a methanogen [12]. Reducing equivalents can be disposed with the production of hydrogen, as long as the methanogen maintains hydrogen concentrations low enough for hydrogen production to be thermodynamically favorable. P. propionicus does not grow in this manner [12]. P. carbinolicus has two highly-similar, four-subunit hydrogenases (Pcar_1602–Pcar_1605 and Pcar_1633–Pcar_1636) related to the cytoplasmic, hydrogen-producing HndABCD hydrogenase in Desulfovibrio fructosovorans [38]. P. propionicus, encodes two highly-similar, 12-subunit enzymes (Ppro_0587–Ppro_0598 and Ppro_3521–Ppro_3532) that are related to the membrane-bound, hydrogen-producing, NiFe hydrogenase in Pyrococcus furiosus [39].

Approximately -45 kJ/mol is predicted to be available to P. carbinolicus and its methanogenic partner when they grow by ethanol oxidation under syntrophic conditions [13]. This is less than what is predicted to be required for generation of an ATP by substrate-level phosphorylation, and it is assumed that some of the ATP from acetyl-CoA conversion must be reinvested to drive the overall reaction, possibly by reverse electron transport [13]. It has been suggested that a Rnf-type ion-translocating electron transfer complex may play a role in reverse electron transport in syntrophs [40], by generating low potential ferredoxin from NADH using the membrane potential [41]. P. carbinolicus contains an Rnf gene cluster (Pcar_0260–Pcar_0265, Figure 4) that is similar to the enzymes from Rhodobacter and Clostridium species, both predicted to be involved in reduction of ferredoxin by NADH [42]. P. propionicus, which does not grow syntrophically, contains no Rnf-like genes.

Both Pelobacter genomes contain unusual numbers and types of ATP synthases, which have been suggested to be important for maintenance of the membrane potential [40]. Both contain the ATP synthase that is conserved in all of the Geobacteraceae, but these genes appear to have recently been duplicated in both organisms (Pcar_0015–Pcar_0016/Pcar_3130–Pcar_3136, Pcar_0944–Pcar_0952, and Ppro_0599–Ppro_0607, Ppro_1500–Ppro_1508, Figures 3 and 4). In addition, both Pelobacter species also have a third ATP synthase encoded in operons organized like that of Methanosarcina barkeri [43], though they were predicted to have been acquired from different sources. The genes in P. propionicus (Ppro_844–Ppro851) are most closely related to enzymes from species of gamma Proteobacteria: Hahella, Legionella, and Methylococcus. The P. carbinolicus genes (Pcar_2989–Pcar2997) are only distantly related to the Geobacter and Desulfuromonas operons, but because they are similar to genes from a variety of species including D. acetoxidans, it is difficult to be confident about the lateral transfer source.

Conclusion

These results provide insights into the evolution of Geobacteraceae species into different environmental niches and biotechnological applications. The results suggest that the last common ancestor of the Geobacteraceae was an acetate-oxidizing, respiratory species capable of extracellular electron transfer, and that specialization for fermentative/syntrophic growth evolved at least twice, allowing some Geobacteraceae to fill additional niches.

The primacy of the respiratory mode is evident from the conservation of genes for all steps in this process including acetate uptake, central metabolism, and electron transfer across the cell membranes in both of the clades of the family. The fermentative/syntrophic Pelobacter species also contain many of these genes. However, they have lost several key enzymes that leave the pathways incomplete, including several necessary for the oxidation of acetate and most of the cytochromes predicted to provide the electrical connection between the inner membrane and the outside of the cell. Instead, the Pelobacter species have appropriated genes via lateral gene transfer for fermentative/syntrophic growth. It is clear that this has happened on two separate occasions. Although both P. carbinolicus and P. propionicus have closely related dehydrogenase genes for the initial metabolism of their unique substrates, acetoin and 2,3-butanediol, the genes for the further fermentation of these substrates are unrelated in the two organisms, reflecting the separate evolution of distinct metabolic pathways. The fact that P. carbinolicus also fills a syntrophic niche, participating in interspecies hydrogen transfer with hydrogen-consuming methanogens, whereas P. propionicus does not, may be explained by the genes associated with reverse electron transfer that only P. carbinolicus has appropriated. The fact that both Pelobacter species have phylogenetically distinct hydrogenase genes that are different from each other as well as those of the Geobacter species, may also reflect the difference in syntrophic capabilities of these species.

The selective pressure to specialize in syntrophic/fermentative growth may have initially been found at the interface of redox boundaries in sedimentary environments. As respiratory Geobacteraceae deplete the supply of the terminal electron acceptor Fe(III) oxide, their capacity for growth is greatly diminished and organisms with other respiratory processes, such as sulfate reduction or methane production, become predominant [2]. Some Geobacter species can oxidize acetate to carbon dioxide and hydrogen when they lack external electron acceptors [44], but the slow rate of this metabolism and the requirement for very low hydrogen partial pressures means that they are not competitive with acetate-utilizing sulfate reducers or methanogens. Acquiring the ability to ferment novel substrates and/or to grow syntrophically could have facilitated expansion into Fe(III) oxide depleted environments. Under such conditions investing energy in the production of respiratory enzymes such as the c-type cytochromes Geobacteraceae require to grow under Fe(III)-reducing conditions would be maladaptive.

These findings also provide insight into the types of metabolic changes that might take place as these organisms are being adapted for modern biotechnical applications. In applications such as the in situ bioremediation of uranium-contaminated groundwater and the conversion of organic compounds to electricity Geobacter species must deal with a scarcity of electron acceptor because electron donor is generally provided well in excess of electron acceptor availability. During in situ bioremediation Fe(III) oxides are rapidly depleted near the source of subsurface acetate amendments, limiting the growth and effectiveness of Geobacter-catalyzed U(VI) reduction [8]. Geobacter species form thick biofilms on the electrodes of microbial fuel cells, forcing many of the cells to metabolize acetate at a significant distance from this artificial electron acceptor [45, 46]. Preliminary results suggest that, like the Pelobacter species described here, Geobacteraceae that predominate during in situ uranium reduction or on the anodes of microbial fuel cells have fewer c-type cytochromes (DRL, unpublished data). Enhanced ability to release excess electrons as hydrogen, in a manner similar to that of P. carbinolicus, could also be beneficial under conditions in which electron acceptor availability is limiting. Thus, these relatively few changes appear to have allowed a respiratory ancestor to radiate out into fermentative and syntrophic niches in addition to their respiratory roles in anaerobic environments. This information serves as a guide to the history of these organisms and provides information that could aid in optimizing their biotechnological applications.

Methods

Genome sequencing and annotation

With the exception of G. sulfurreducens [16], sequence data for the Geobacteraceae genomes were produced by the US Department of Energy Joint Genome Institute http://www.jgi.doe.gov, using a whole-genome shotgun strategy for the Sanger-sequencing of 3-Kb, 8-Kb, and 40-Kb DNA libraries to 8-9X depth. Open reading frames and their translations and predicted function based on automated annotation were taken from NCBI http://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/ (Table 1).

Clustering orthologs into protein families

All proteins in the genomes were clustered into families of orthologs and recent paralogs using OrthoMCL [17], which uses reciprocal best similarity pairs from all-vs-all BLAST [47] to identify orthologs and recent paralogs, which are then clustered together across all the genomes using the Markov clustering algorithm [48]. A functional role was predicted for each cluster using the G. sulfurreducens in silico model annotation [18] and COG categorization [19].

Phylogenomics

All the ORFs from the six Geobacteraceae genomes and the outgroup species Anaeromyxobacter dehalogenans 2CP-C (NC_007760) were put into orthologous groups using Hal [49], with inflation parameters from 1.1–5.0 for the clustering algorithm. The proteins used for the phylogeny were those that were part of a cluster generated with any inflation value that had exactly one member from each genome (see Additional file 6). All of the proteins in the cluster were concatenated and the resulting sequences aligned by ClustalW [50]. ProtTest [51] was used to select a model of molecular evolution and MrBayes [52] was used to create a Bayesian estimation of the phylogeny.

Lateral gene transfer

A phylogenetic tree was inferred using PhyloGenie [53] for every protein from the six genomes. Homologous sequences for each protein were selected by BLAST from the non-redundant protein database from NCBI http://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/, alignments were created with ClustalW [50], and the phylogeny was inferred using neighbor-joining [54] and 100 bootstrapped replicates. These trees were used to identify proteins for which the nearest relative was not from the Geobacteraceae. If the phylogeny was strongly supported (bootstrap ≥ 50) or if the phylogeny was weakly supported and the most similar sequence in the non-redundant protein database from NCBI was not a Geobacteraceae species, the protein was considered a candidate. If the next branch out contained a single sequence not from Geobacteraceae species, the query gene was defined as being from lateral transfer. If the next branch contained a single sequence from Geobacteraceae, it was not. If the sister group was a clade or was not strongly supported, the ancestral condition was inferred [55] and used to determine lateral transfer.

References

  1. Lovley DR, Holmes DE, Nevin KP: Dissimilatory Fe(III) and Mn(IV) reduction. Adv Microb Physiol. 2004, 49: 219-286. 10.1016/S0065-2911(04)49005-5.

    Article  CAS  PubMed  Google Scholar 

  2. Lovley DR: Dissimilatory Fe(III) and Mn(IV) reduction. Microbiol Rev. 1991, 55 (2): 259-287.

    PubMed Central  CAS  PubMed  Google Scholar 

  3. Lovley DR: Cleaning up with genomics: applying molecular biology to bioremediation. Nat Rev Microbiol. 2003, 1 (1): 35-44. 10.1038/nrmicro731.

    Article  CAS  PubMed  Google Scholar 

  4. Lovley DR, Nevin KP: Electricity production with electricigens. Bioenergy. Edited by: Wall J. 2008, Washington, DC: ASM Press, 295-306.

    Chapter  Google Scholar 

  5. Lovley DR, Coates JD, Blunt-Harris EL, Phillips EJ, Woodward JC: Humic substances as electron acceptors for microbial respiration. Nature. 1996, 382: 445-448. 10.1038/382445a0.

    Article  CAS  Google Scholar 

  6. Lovley DR, Woodward JC, Chapelle FH: Stimulated anoxic biodegradation of aromatic hydrocarbons using Fe(III) ligands. Nature. 1994, 370 (6485): 128-131. 10.1038/370128a0.

    Article  CAS  PubMed  Google Scholar 

  7. Lovley DR, Phillips EJ, Gorby YA, Landa ER: Microbial reduction of uranium. Nature. 1991, 350: 413-416. 10.1038/350413a0.

    Article  CAS  Google Scholar 

  8. Anderson RT, Vrionis HA, Ortiz-Bernad I, Resch CT, Long PE, Dayvault R, Karp K, Marutzky S, Metzler DR, Peacock A, White DC, Lowe M, Lovley DR: Stimulating the in situ activity of Geobacter species to remove uranium from the groundwater of a uranium-contaminated aquifer. Appl Environ Microbiol. 2003, 69 (10): 5884-5891. 10.1128/AEM.69.10.5884-5891.2003.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  9. Bond DR, Holmes DE, Tender LM, Lovley DR: Electrode-reducing microorganisms that harvest energy from marine sediments. Science. 2002, 295 (5554): 483-485. 10.1126/science.1066771.

    Article  CAS  PubMed  Google Scholar 

  10. Lovley DR: Bug juice: harvesting electricity with microorganisms. Nat Rev Microbiol. 2006, 4 (7): 497-508. 10.1038/nrmicro1442.

    Article  CAS  PubMed  Google Scholar 

  11. Tender LM, Reimers CE, Stecher HA, Holmes DE, Bond DR, Lowy DA, Pilobello K, Fertig SJ, Lovley DR: Harnessing microbially generated power on the seafloor. Nat Biotechnol. 2002, 20 (8): 821-825.

    Article  CAS  PubMed  Google Scholar 

  12. Schink B: Fermentation of 2,3-butanediol by Pelobacter carbinolicus sp. nov. and Pelobacter propionicus sp. nov., and evidence for propionate formation from C-2 compounds. Arch Microbiol. 1984, 137 (1): 33-41. 10.1007/BF00425804.

    Article  CAS  Google Scholar 

  13. Schink B: Energetics of syntrophic cooperation in methanogenic degradation. Microbiol Mol Biol Rev. 1997, 61 (2): 262-280.

    PubMed Central  CAS  PubMed  Google Scholar 

  14. Haveman SA, Didonato RJ, Villanueva L, Shelobolina ES, Postier B, Xu A, Liu A, Lovley DR: Genome-Wide Gene Expression Patterns and Growth Requirements Suggest that Pelobacter carbinolicus Reduces Fe(III) Indirectly via Sulfide Production. Appl Environ Microbiol. 2008,

    Google Scholar 

  15. Richter H, Lanthier M, Nevin KP, Lovley DR: Lack of electricity production by Pelobacter carbinolicus indicates that the capacity for Fe(III) oxide reduction does not necessarily confer electron transfer ability to fuel cell anodes. Appl Environ Microbiol. 2007, 73 (16): 5347-5353. 10.1128/AEM.00804-07.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  16. Methe BA, Nelson KE, Eisen JA, Paulsen IT, Nelson W, Heidelberg JF, Wu D, Wu M, Ward N, Beanan MJ, Dodson RJ, Madupu R, Brinkac LM, Daugherty SC, DeBoy RT, Durkin AS, Gwinn M, Kolonay JF, Sullivan SA, Haft DH, Selengut J, Davidsen TM, Zafar N, White O, Tran B, Romero C, Forberger HA, Weidman J, Khouri H, Feldblyum TV, et al: Genome of Geobacter sulfurreducens: metal reduction in subsurface environments. Science. 2003, 302 (5652): 1967-1969. 10.1126/science.1088727.

    Article  CAS  PubMed  Google Scholar 

  17. Li L, Stoeckert CJ, Roos DS: OrthoMCL: identification of ortholog groups for eukaryotic genomes. Genome Res. 2003, 13 (9): 2178-2189. 10.1101/gr.1224503.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  18. Mahadevan R, Bond DR, Butler JE, Esteve-Nunez A, Coppi MV, Palsson BO, Schilling CH, Lovley DR: Characterization of metabolism in the Fe(III)-reducing organism Geobacter sulfurreducens by constraint-based modeling. Appl Environ Microbiol. 2006, 72 (2): 1558-1568. 10.1128/AEM.72.2.1558-1568.2006.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  19. Tatusov RL, Koonin EV, Lipman DJ: A genomic perspective on protein families. Science. 1997, 278 (5338): 631-637. 10.1126/science.278.5338.631.

    Article  CAS  PubMed  Google Scholar 

  20. Holmes DE, Nevin KP, Lovley DR: Comparison of 16S rRNA, nifD, recA, gyrB, rpoB and fusA genes within the family Geobacteraceae fam. nov. Int J Syst Evol Microbiol. 2004, 54 (Pt 5): 1591-1599. 10.1099/ijs.0.02958-0.

    Article  CAS  PubMed  Google Scholar 

  21. Champine JE, Goodwin S: Acetate catabolism in the dissimilatory iron-reducing isolate GS-15. J Bacteriol. 1991, 173 (8): 2704-2706.

    PubMed Central  CAS  PubMed  Google Scholar 

  22. Galushko AS, Schink B: Oxidation of acetate through reactions of the citric acid cycle by Geobacter sulfurreducens in pure culture and in syntrophic coculture. Arch Microbiol. 2000, 174 (5): 314-321. 10.1007/s002030000208.

    Article  CAS  PubMed  Google Scholar 

  23. Paulsen J, Kroger A, Thauer RK: ATP-driven succinate oxidation in the catabolism of Desulfuromonas acetoxidans. Arch Microbiol. 1986, 144: 202-207. 10.1007/BF00454960.

    Article  Google Scholar 

  24. Risso C, Methe BA, Elifantz H, Holmes DE, Lovley DR: Highly Conserved Genes in Geobacter Species with Expression Patterns Indicative of Acetate Limitation. Microbiology. 2008,

    Google Scholar 

  25. Imabayashi F, Aich S, Prasad L, Delbaere LT: Substrate-free structure of a monomeric NADP isocitrate dehydrogenase: an open conformation phylogenetic relationship of isocitrate dehydrogenase. Proteins. 2006, 63 (1): 100-112. 10.1002/prot.20867.

    Article  CAS  PubMed  Google Scholar 

  26. Coppi MV, O'Neil RA, Leang C, Kaufmann F, Methe BA, Nevin KP, Woodard TL, Liu A, Lovley DR: Involvement of Geobacter sulfurreducens SfrAB in acetate metabolism rather than intracellular, respiration-linked Fe(III)-citrate reduction. Microbiology. 2007, 153 (10): 3572-3585. 10.1099/mic.0.2007/006478-0.

    Article  CAS  PubMed  Google Scholar 

  27. Tseng CP, Yu CC, Lin HH, Chang CY, Kuo JT: Oxygen- and growth rate-dependent regulation of Escherichia coli fumarase (FumA, FumB, and FumC) activity. J Bacteriol. 2001, 183 (2): 461-467. 10.1128/JB.183.2.461-467.2001.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  28. Schutz M, Brugna M, Lebrun E, Baymann F, Huber R, Stetter KO, Hauska G, Toci R, Lemesle-Meunier D, Tron P, Schmidt C, Nitschke W: Early evolution of cytochrome bc complexes. J Mol Biol. 2000, 300 (4): 663-675. 10.1006/jmbi.2000.3915.

    Article  CAS  PubMed  Google Scholar 

  29. Allen JW, Daltrop O, Stevens JM, Ferguson SJ: C-type cytochromes: diverse structures and biogenesis systems pose evolutionary problems. Philos Trans R Soc Lond B Biol Sci. 2003, 358 (1429): 255-266. 10.1098/rstb.2002.1192.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  30. Butler JE, Kaufmann F, Coppi MV, Nunez C, Lovley DR: MacA, a diheme c-type cytochrome involved in Fe(III) reduction by Geobacter sulfurreducens. J Bacteriol. 2004, 186 (12): 4042-4045. 10.1128/JB.186.12.4042-4045.2004.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  31. Leang C, Coppi MV, Lovley DR: OmcB, a c-type polyheme cytochrome, involved in Fe(III) reduction in Geobacter sulfurreducens. J Bacteriol. 2003, 185 (7): 2096-2103. 10.1128/JB.185.7.2096-2103.2003.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  32. Oppermann FB, Steinbuchel A: Identification and molecular characterization of the aco genes encoding the Pelobacter carbinolicus acetoin dehydrogenase enzyme system. J Bacteriol. 1994, 176 (2): 469-485.

    PubMed Central  CAS  PubMed  Google Scholar 

  33. Huang M, Oppermann FB, Steinbuchel A: Molecular characterization of the Pseudomonas putida 2,3-butanediol catabolic pathway. FEMS Microbiol Lett. 1994, 124 (2): 141-150. 10.1111/j.1574-6968.1994.tb07276.x.

    Article  CAS  PubMed  Google Scholar 

  34. Schink B, Kremer DR, Hansen TA: Pathway of propionate formation from ethanol in Pelobacter propionicus. Arch Microbiol. 1987, 147: 321-327. 10.1007/BF00406127.

    Article  CAS  Google Scholar 

  35. Reddy DV, Shenoy BC, Carey PR, Sonnichsen FD: High resolution solution structure of the 1.3S subunit of transcarboxylase from Propionibacterium shermanii. Biochemistry. 2000, 39 (10): 2509-2516. 10.1021/bi9925367.

    Article  CAS  PubMed  Google Scholar 

  36. Butler JE, Glaven RH, Esteve-Nunez A, Nunez C, Shelobolina ES, Bond DR, Lovley DR: Genetic characterization of a single bifunctional enzyme for fumarate reduction and succinate oxidation in Geobacter sulfurreducens and engineering of fumarate reduction in Geobacter metallireducens. J Bacteriol. 2006, 188 (2): 450-455. 10.1128/JB.188.2.450-455.2006.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  37. Coppi MV: The hydrogenases of Geobacter sulfurreducens: a comparative genomic perspective. Microbiology. 2005, 151 (Pt 4): 1239-1254. 10.1099/mic.0.27535-0.

    Article  CAS  PubMed  Google Scholar 

  38. Malki S, Saimmaime I, De Luca G, Rousset M, Dermoun Z, Belaich JP: Characterization of an operon encoding an NADP-reducing hydrogenase in Desulfovibrio fructosovorans. J Bacteriol. 1995, 177 (10): 2628-2636.

    PubMed Central  CAS  PubMed  Google Scholar 

  39. Sapra R, Verhagen MF, Adams MW: Purification and characterization of a membrane-bound hydrogenase from the hyperthermophilic archaeon Pyrococcus furiosus. J Bacteriol. 2000, 182 (12): 3423-3428. 10.1128/JB.182.12.3423-3428.2000.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  40. McInerney MJ, Rohlin L, Mouttaki H, Kim U, Krupp RS, Rios-Hernandez L, Sieber J, Struchtemeyer CG, Bhattacharyya A, Campbell JW, Gunsalus RP: The genome of Syntrophus aciditrophicus: life at the thermodynamic limit of microbial growth. Proc Natl Acad Sci USA. 2007, 104 (18): 7600-7605. 10.1073/pnas.0610456104.

    Article  PubMed Central  PubMed  Google Scholar 

  41. Kumagai H, Fujiwara T, Matsubara H, Saeki K: Membrane localization, topology, and mutual stabilization of the rnfABC gene products in Rhodobacter capsulatus and implications for a new family of energy-coupling NADH oxidoreductases. Biochemistry. 1997, 36 (18): 5509-5521. 10.1021/bi970014q.

    Article  CAS  PubMed  Google Scholar 

  42. Kim J, Hetzel M, Boiangiu CD, Buckel W: Dehydration of (R)-2-hydroxyacyl-CoA to enoyl-CoA in the fermentation of alpha-amino acids by anaerobic bacteria. FEMS Microbiol Rev. 2004, 28 (4): 455-468.

    Article  CAS  PubMed  Google Scholar 

  43. Sumi M, Yohda M, Koga Y, Yoshida M: F0F1-ATPase genes from an archaebacterium, Methanosarcina barkeri. Biochem Biophys Res Commun. 1997, 241 (2): 427-433. 10.1006/bbrc.1997.7809.

    Article  CAS  PubMed  Google Scholar 

  44. Cord-Ruwisch R, Lovley DR, Schink B: Growth of Geobacter sulfurreducens with acetate in syntrophic cooperation with hydrogen-oxidizing anaerobic partners. Appl Environ Microbiol. 1998, 64 (6): 2232-2236.

    PubMed Central  CAS  PubMed  Google Scholar 

  45. Nevin KP, Richter H, Covalla SF, Johnson JP, Woodard TL, Orloff AL, Jia H, Zhang M, Lovley DR: Power output and columbic efficiencies from biofilms of Geobacter sulfurreducens comparable to mixed community microbial fuel cells. Environ Microbiol. 2008

    Google Scholar 

  46. Reguera G, Nevin KP, Nicoll JS, Covalla SF, Woodard TL, Lovley DR: Biofilm and nanowire production leads to increased current in Geobacter sulfurreducens fuel cells. Appl Environ Microbiol. 2006, 72 (11): 7345-7348. 10.1128/AEM.01444-06.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  47. Altschul SF, Madden TL, Schaffer AA, Zhang J, Zhang Z, Miller W, Lipman DJ: Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997, 25 (17): 3389-3402. 10.1093/nar/25.17.3389.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  48. Enright AJ, Van Dongen S, Ouzounis CA: An efficient algorithm for large-scale detection of protein families. Nucleic Acids Res. 2002, 30 (7): 1575-1584. 10.1093/nar/30.7.1575.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  49. Robbertse B, Reeves JB, Schoch CL, Spatafora JW: A phylogenomic analysis of the Ascomycota. Fungal Genet Biol. 2006, 43 (10): 715-725. 10.1016/j.fgb.2006.05.001.

    Article  CAS  PubMed  Google Scholar 

  50. Thompson JD, Higgins DG, Gibson TJ: CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res. 1994, 22 (22): 4673-4680. 10.1093/nar/22.22.4673.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  51. Abascal F, Zardoya R, Posada D: ProtTest: selection of best-fit models of protein evolution. Bioinformatics. 2005, 21 (9): 2104-2105. 10.1093/bioinformatics/bti263.

    Article  CAS  PubMed  Google Scholar 

  52. Ronquist F, Huelsenbeck JP: MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics. 2003, 19 (12): 1572-1574. 10.1093/bioinformatics/btg180.

    Article  CAS  PubMed  Google Scholar 

  53. Frickey T, Lupas AN: PhyloGenie: automated phylome generation and analysis. Nucleic Acids Res. 2004, 32 (17): 5231-5238. 10.1093/nar/gkh867.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  54. Saitou N, Nei M: The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol. 1987, 4 (4): 406-425.

    CAS  PubMed  Google Scholar 

  55. Swofford DL, Madison WP: Reconstructing ancestral character states under Wagner parsimony. Math Biosci. 1987, 87: 199-229. 10.1016/0025-5564(87)90074-5.

    Article  Google Scholar 

Download references

Acknowledgements

We thank Peter Brown for his assistance. This research was supported by the Office of Science (BER), U. S. Department of Energy, Cooperative Agreement No. DE-FC02-02ER63446.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jessica E Butler.

Additional information

Authors' contributions

JEB carried out the conservation analysis, created the physiological models, constructed the single gene phylogenies, and drafted the manuscript. NDY designed the method for and carried out the lateral gene transfer prediction, constructed the whole genome phylogeny, and carried out the clustering method. DRL conceived of the study and helped draft the manuscript. All authors read and approved the final manuscript.

Electronic supplementary material

12864_2008_1987_MOESM1_ESM.xls

Additional file 1:All proteins referenced in this study. Spreadsheet with NCBI identification numbers and descriptions including name, predicted function, COG membership, size, location in genome, protein family, family size, family conservation pattern, lateral transfer prediction, and heme-binding motifs.(XLS 6 MB)

12864_2008_1987_MOESM2_ESM.xls

Additional file 2:Protein families and their members from each of the genomes. Spreadsheet showing protein families of orthologs, with descriptions including member proteins, member genomes, family size, and conservation pattern.(XLS 585 KB)

12864_2008_1987_MOESM3_ESM.xls

Additional file 3: Frequency of the phyletic patterns of protein conservation. Spreadsheet showing the number of proteins with a given pattern of conservation. (XLS 20 KB)

12864_2008_1987_MOESM4_ESM.xls

Additional file 4: Proteins conserved only within the respiratory species. Spreadsheet showing the proteins that were conserved in all and only the respiratory species. (XLS 124 KB)

12864_2008_1987_MOESM5_ESM.xls

Additional file 5: Proteins conserved only within the fermentative species. Spreadsheet showing the proteins that were conserved in all and only the fermentative species. (XLS 94 KB)

12864_2008_1987_MOESM6_ESM.xls

Additional file 6: Proteins used in the whole genome phylogeny. Spreadsheet showing the proteins used and their COG single-letter code and COG functional categorization. (XLS 328 KB)

12864_2008_1987_MOESM7_ESM.xls

Additional file 7: Conservation patterns for proteins involved in the energy metabolism of anaerobic respiration. Spreadsheet showing all of the proteins with their metabolic role, conservation pattern, reaction abbreviation in the constraint-based model, and protein family. (XLS 28 KB)

12864_2008_1987_MOESM8_ESM.ppt

Additional file 8: Phylogeny of the Geobacteraceae isocitrate dehydrogenases. Figure showing a neighbor joining model of the phylogeny of these proteins with NCBI identification numbers. (PPT 101 KB)

12864_2008_1987_MOESM9_ESM.xls

Additional file 9: Genes predicted to have been acquired by lateral gene transfer. Spreadsheet showing only those genes predicted to have been acquired by lateral gene transfer. (XLS 761 KB)

12864_2008_1987_MOESM10_ESM.ppt

Additional file 10: Phylogeny of the Pelobacter butanediol dehydrogenases. Figure showing a neighbor joining model of the phylogeny of these proteins with NCBI identification numbers. (PPT 42 KB)

Authors’ original submitted files for images

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/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Butler, J.E., Young, N.D. & Lovley, D.R. Evolution from a respiratory ancestor to fill syntrophic and fermentative niches: comparative genomics of six Geobacteraceae species. BMC Genomics 10, 103 (2009). https://0-doi-org.brum.beds.ac.uk/10.1186/1471-2164-10-103

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/1471-2164-10-103

Keywords