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

Identification of constraints influencing the bacterial genomes evolution in the PVC super-phylum

Abstract

Background

Horizontal transfer plays an important role in the evolution of bacterial genomes, yet it obeys several constraints, including the ecological opportunity to meet other organisms, the presence of transfer systems, and the fitness of the transferred genes. Bacteria from the Planctomyctetes, Verrumicrobia, Chlamydiae (PVC) super-phylum have a compartmentalized cell plan delimited by an intracytoplasmic membrane that might constitute an additional constraint with particular impact on bacterial evolution. In this investigation, we studied the evolution of 33 genomes from PVC species and focused on the rate and the nature of horizontally transferred sequences in relation to their habitat and their cell plan.

Results

Using a comparative phylogenomic approach, we showed that habitat influences the evolution of the bacterial genome’s content and the flux of horizontal transfer of DNA (HT). Thus bacteria from soil, from insects and ubiquitous bacteria presented the highest average of horizontal transfer compared to bacteria living in water, extracellular bacteria in vertebrates, bacteria from amoeba and intracellular bacteria in vertebrates (with a mean of 379 versus 110 events per species, respectively and 7.6% of each genomes due to HT against 4.8%). The partners of these transfers were mainly bacterial organisms (94.9%); they allowed us to differentiate environmental bacteria, which exchanged more with Proteobacteria, and bacteria from vertebrates, which exchanged more with Firmicutes. The functional analysis of the horizontal transfers revealed a convergent evolution, with an over-representation of genes encoding for membrane biogenesis and lipid metabolism, among compartmentalized bacteria in the different habitats.

Conclusions

The presence of an intracytoplasmic membrane in PVC species seems to affect the genome’s evolution through the selection of transferred DNA, according to their encoded functions.

Background

The extensive amount of genomic data acquired over the last 20 years has provided insights into the evolutionary processes that drive bacterial evolution. The horizontal transfer of DNA (HT) appears to be major driving force of innovation [1, 2] as it provides additional functions, allowing adaptation to specific conditions and environmental changes. The HT process in bacteria depends on several conditions [3]: i. the possibility of exchanges, meaning the presence of different microorganisms in a single place; ii. the possibility of foreign sequences to enter into recipient bacteria, mediated by conjugation, transformation or transduction; iii. the ability to integrate into the recipient genome; iv. the genes expressed and the genes used v. those conserved, in relation to the benefits for recipient bacteria. This process could be regulated by intrinsic and extrinsic constraints. Two extrinsic constraints influencing the possibility of exchanges include the environment or the “ecological niches” and the lifestyle, which together constitute the habitat of bacteria [46]. Thus the proportions and origins of HT were more similar among bacteria from the same habitat than among bacteria from a given phylum [7]. Changing environmental conditions are also well known constraints for HT regulation; UV irradiation or starvation and other stress conditions, were shown to affect the mobility of transposons and insertion sequences [6, 810]. The habitat also seems to play an important role in the selection and conservation of transferred sequences encoding for specific functions that are involved in host’s colonization, and the development of pathogenesis. Indeed, many examples in the literature indicate that genes encoding for metabolic functions [1113] and for antibiotic resistance [14, 15] and virulence [1618] represent commonly transferred sequences. The intrinsic constraints that influence the entrance and integration of foreign DNA into a recipient genome include the exclusion surface that limits the entrance of specific sequences in some bacteria [3], the presence of CRISPR that decreases the quantity of transferred sequences insertion in recipient genomes [19, 20] and the presence of some endo-nucleases that can destroy foreign DNA [3, 21].

Many studies have been conducted to explore the impact of the different extrinsic and intrinsic constraints on horizontal transfers. However, these studies involved one or a few species, or bacteria presenting only one to two habitats or lifestyles [2225], or undergoing relatively few intrinsic constraints [26, 27]. The study of only few characteristics may one lead to miss the cumulative or overlapping effects of the different constraints. Therefore, we used a phylogenomic approach to mine a large set of bacteria with different habitats in order to decipher the impact of different constraints on genome composition, especially regarding HT. The PVC super-phylum seems to be a good model to study, as it includes seven bacterial phyla (Planctomycetes, Verrucomicrobiae, Chlamydiae, Lentisphaera, Poribacteria, OP3, WWE2) [2831] with diverse habitats, three different lifestyles (intracellular allopatric, intracellular sympatric, extracellular sympatric) and numerous environments (water, soils, water and soils, metazoa, amoeba, ubiquitous…), thus varying the external constraints. Moreover, a specific cell plan is also present in all the Planctomycetes [3234], in some of Verrucomicrobiae [35] in one Lentisphaera and in one Poribacteria [36]. The cytoplasm of these bacteria is separated into two compartments by an intracytoplasmic membrane (ICM), the pirellulosome inside (with DNA [37]) and the paryphoplasm outside. This membrane is a lipid bilayer in contact with proteins [32, 33, 38] presenting structural similarities with proteins from eukaryotic membranes like the clathrins [39, 40]. The function of this intracytoplasmic membrane is still unknown, but we hypothesize the possible impact of this intrinsic constraint on HT. In the present investigation, we analyzed 33 PVC bacteria together with 31 phylogenetically close species (Bacteroidetes, Chlorobi and Spirochaetes) that were considered as the control group, looking for evidence for horizontal transfer. Statistical analyses of the potential partners and functions involved in HT allowed us to estimate the real impact of habitat and cell plan on the genomes evolution.

Methods

Bacterial set selection, definition of lifestyles, environments and cell plan

The genomes of 64 bacteria have been retrieved from two different databases [41, 42]. These bacteria belong to different phyla (Additional file 1) including four phyla of the PVC super-phylum, Planctomycetes, Verrucomicrobiae, Lentisphaerae and Chlamydia and the phylogenetically closest phyla, Bacteroidetes, Chlorobi and Spirochaeta (determined thanks to a reference tree [43]). We reconstructed the species tree of PVC bacteria and Bacteroidetes-Chlorobi-Spirochaetes on the basis of 12 markers that are common to the 64 species (Additional file 2) using Mega5 [44]. Therefore, the protein sequences of each marker were aligned with Muscle [45] and non-conserved positions were removed manually. All alignments were concatenated, leading to an alignment of 5067 sites. We used a Maximum likelihood tree (substitution model JTT) based on this concatenated alignment to reconstruct the phylogeny of species. Bootstrap support values were obtained with 150 replicates (Additional file 3). The bacteria studied have different lifestyles (intracellular or extracellular, allopatric or sympatric [46, 47]) and live in different environments (amoeba, mammals, soils, water, insects). We use the term “environment” as the main place where the bacteria are living. For example, bacteria detected in sea, freshwater or wastewater are all annotated as bacteria from ‘water’. If bacteria are present in two different environments, we indicate both of them (for example ‘water-soil’ bacteria), while bacteria living in more than 5 environments are considered ‘ubiquitous’. The lifestyle of bacteria is characterized by two factors: the intracellular and extracellular conditions, and the ability to exchange with other microorganisms (in allopatric or sympatric lifestyles, respectively [46]). Lifestyles and living environments were defined for each bacterium based on a literature search [4853]. Cell plans of the bacteria were determined via transmission electron microscopy images already available in the literature [3336] and microscopic observations of the bacteria realized in our laboratory [54]. Three states are determined for the cell plan: compartmentalization, non compartmentalization and unknown. The selected set of species contains 4 bacteria from amoeba, 4 from insects (1 intracellular, 3 extracellular), 3 from soils, 4 living in soils and water, 3 ubiquitous, 26 from vertebrates (18 extracellular, 8 intracellular) and 20 from water. Among these 64 bacteria, 20 present a compartmentalized cell plan, 5 have an unknown cell plan, and 39 are not compartmentalized (Additional file 1).

Genome analysis: common genes, specific genes, ORFans

OrthoMCL [55] was used to obtain groups of orthologous proteins. Groups containing at least one representative member of each habitat were considered as the common genes, and those that contain only proteins of bacteria from the same habitat are considered as specific to the corresponding habitat. We calculated the rate and determined the function of proteins that are specific to habitat in each species using two software, COGnitor and WGA and the Interpro database [5659]. Genes that do not belong to any orthologous group are either acquired by “specific” HT or generated de novo. Blast against NR database allowed the identification of ORFans in the genome with no identifiable homologous (i.e. genes that do not have a Blast hit with an e-value < 10e-4 AND a query coverage >50%). We performed a clustering of all species according to their genes contents in order to detect, in some bacteria, a tendency to share the same gene contents in relation with their habitat.

HT detection, functions and partner identification

Generally, two main difficulties hinder the analysis of the Horizontal Transfer of DNA sequences (HT) according to habitat: the distinction between the ancestral and recent gene gains, as well as the difficulty of determining the ancestral habitat of the bacteria. In order to avoid these problems, we focused our study on recent transfers that occurred only in modern species of the super-phylum (in the leafs of the tree) and not in their ancestors (at the nodes of the tree). HT instances were identified using a comparative phylogenomic approach, phylogenetic profiling of proteins and phylogenetic analysis of gene trees in comparison with the species tree. Using the Phylopattern [60] pipeline, we identified the gene gain events in four steps: 1) based on the orthologous groups, a tree was reconstructed for each group. 2) The topologies of these trees were compared with that of the species tree in order to detect species missing in orthologous groups. 3) We obtained a pattern of presence/absence for each gene in the different species, allowing Phylopattern to reconstruct the ancestral states of genes by implementing the Sankoff parsimony algorithm [61]. Based on this reconstruction, the pattern-matching module in PhyloPattern allows us to infer, by parsimony, two types of genetic events that could have occurred during gene evolution: gains and losses. The gains could be a possible HT, de novo genes or artifacts, therefore, among gene gains detected by Phylopattern, we had to identify those due to HT. We focused on specific gene gains and performed a Blast to identify similar sequences in the NR database. If the first twenty hits of Blast result belong to species outside the super-phylum of the query, and they are orthologous as confirmed by a reciprocal best hit, the enquired gene could be considered horizontally acquired. The pattern permitting the automatic identification of HT among gain events is presented in Additional file 4. Sequences with e-value > 10–5, coverage < 60% or identities < 30% were not considered. The localization of these events in the genome allowed the identification of any horizontal transfers of DNA sequences. The directionality of the transfer could not always be identified, but as we were interested in the capacity of exchange of the bacteria, it did not matter if the bacterium was the donor or the recipient. If some transferred genes are side by side in genomes, and were exchanged with the same partners, we considered them to have been transferred by a single event.

We calculated quantities and proportions of proteins (proteins transferred/total proteins in proteome) and sequences (nucleotides transferred/total nucleotides in genomes) implicated in HT for each genomes and the size of the transferred sequences. We identified the function of transferred genes by using two software programs, COGnitor and WGA and the Interpro database [5659]). These programs attribute one type of function to proteins, according to the COG to which proteins belong. We studied the possible significant differences in HT distribution among the different studied groups of bacteria concerning the HT partners and functions.

Statistical analyses

The occurrence of HT and the frequency of specific genes were compared among the different groups of bacteria from different habitats. We tested whether the data (proportions of functions for specific genes and proportions of partners and functions for HT) follows a normal (Gaussian) distribution using the Shapiro-Wilk test and we controlled the homogeneity of the data by the Levene test [62]. These tests were followed by a comparison of variance among the different habitats, using the Kruskal test [63] or the ANOVA test, according to whether or not the data had a normal distribution. The Nemenyi [64] or Tukey [65] tests were performed to obtain a comparison of each pair of habitats. We also realized a Principal Components Analysis (PCA), focused on HT proportions in genomes, size, functions and partner of transfer, followed by a hierarchical clustering (HCPC), to identify clustering of bacteria according to their transfer partners or their functions. All analyses were realized with R software.

We then used comparative phylogenetic methods to test the impact of phylogenetic relationships between species on the acquisition of studied characters. Analysis of variance was used in intergroup comparisons of categorical variables to determine whether there is statistical significance of the repartition of species on the basis of their habitat, compared to their classification according to phylogenetic distances (Additional file 3). Therefore, the Nemenyi [64] or Tukey [65] tests were performed to obtain a comparison of each pairs of groups based on the phylogenetic relationships. These groups were determined by the phylogenetic distance separating bacteria. We compared the results of these tests with the results of tests carried out for groups based on habitats, allowing us to determine if classes defined by the phylogenetic distances present different and more reliable results than classes based on habitat (t-test). If results were similar, it could be difficult to determine whether the differences observed were related to the habitat or to the phylogenetic relationships. We also performed two correlation tests. The first test, Pagel’s correlation method [66], performed on Mesquite, is a test of the independent evolution of two binary characters (all features studied were tested thanks to a binarization of continuous values). This test compares the ratio of likelihoods of two models where the rates of change in each character are dependent or alternatively independent from phylogenetic relationships. The second test is the Spearman coefficient [67], weighted by the phylogenetic distances that studies the relationship between two variables. The detection by means of this correlation test of a significantly convergent character in bacteria from a single habitat is rather unrelated to the phylogenetic background.

Results

Pangenome analysis

The genome size varied widely among the 64 bacteria studied, ranging from 0.63 Mb for Blattabacterium to 9.76 Mb for Singulisphaera acidiphila with an average of 3.83 +/− 2.2 Mb. Bacteria from soils, insects and water-soils presented the largest genome sizes, with an average of 7.07, 6.45 and 5.69 Mb, respectively, while the smallest genome sizes were found among intracellular and extracellular bacteria of vertebrates, and bacteria from amoeba, with an average of 1.14, 2.56 and 2.91 Mb, respectively. Ubiquitous bacteria and bacteria from water with an average genome size of 4.63 Mb, formed the medium-sized genomes. Likewise, the protein sets were very different among the studied bacteria, ranging from 579 proteins for Blattabacterium sp to 7969 for Gemmata obscuriglobus, with an average of 3227 +/− 2513. When using OrthoMCL, 124,175 out of 206,508 proteins that form the pangenome of the PVC group bacteria, could be assigned to 16,918 different orthologous groups (OG). Among these, 1224 OGs were common to the eight different habitats studied, and constituted the common genes. The rest of the genes were present in bacteria from two or more habitats, and thus form the “shared genome.” or they did not share sequence similarity with any other gene of the species of other habitats, and thus constituted the “specific genes” (Fig. 1). When species were clustered according to their gene contents, some bacteria sharing a same ecological niche were preferentially grouped together, forming subclusters within the clusters, determined by phylogenetic relationships or disturbing the phylogenetic unity of some groups, like the Verrucomicrobiae, Spirochaetes and Planctomycetes (Additional file 5). Thus the ecological niche seems to have influenced the gene content of some bacteria from soil (p-value = 0.027) and water (p-value = 0.045 for internal cluster). Bacteria from insects that showed the highest proportion of their genes shared with the other groups and the lowest quantity of specific genes (29 genes) were scattered throughout the different clusters (Additional file 5).

Fig. 1
figure 1

Comparison of the genome contents in the different habitats. The quantity of genes common to all habitats (common genes), and their functions, are indicated in the center of the flower. Genes specific to each habitat and genes shared by different habitats are represented by the different petals. The clouds represent the ORFans percentages in genomes. The width of the petals and the clouds is proportional to the quantity of specific genes and percentage of ORFans, respectively. The functional distribution of common, shared and specific genes is presented in the table

The genes that are common to all habitats represented 26.2% of the content of each genome on average, and varied from 20.1% for bacteria from insects to 49.5% for the intracellular vertebrates. In order to determine the functional profile of the common genes, each protein was assigned to Cluster of Orthologous Groups of proteins (COGs) functional category. We could infer a putative function to the protein sequences of 74% of the common genes; of these, 43.8% encode for cellular processes and signaling, 36.8% for metabolic functions, and 19.5% for storage and processing information (Fig. 1). Among these, four functions were significantly over-represented compared to the other functions: wall/membrane/envelop biogenesis, signal transduction mechanisms, transcription and energy production and conversion (12.4, 11.6, 9.1 and 8.5%, respectively Chi2 test: p-value = 9.4*10-7) (Fig. 1).

The genes shared by some habitats and the specific genes represented 35.3 and 8.1% of the content of each genome, respectively on average. The specific genes varied from 0.5% in bacteria from insects to 18.4% in the intracellular vertebrates. Of all species analyzed, Bacteroides xylanisolvens and Bacteroides vulgatus in the group of extracellular vertebrates, had the most specific genes, with a total of 801 and 780 exclusive sequences (18.2 and 19.2% of their total genomes), respectively. The functional distribution of the specific genes was significantly different from that of the common genes in all the habitats with fewer genes implicated in cell process and signaling (37.9%) (t-test for comparison between specific genes and common genes; p-value = 4.3*10-4) (Fig. 1). Some functions were significantly over-represented in some habitats compared to other habitats (Fig. 2), including transcription within their specific genes (16.3 and 18.8%, Kruskal-Wallis test: p-value = 3.9*10-6 and Correlation test : p-value = 4.7*10-3) in bacteria from amoeba and from soils-water; the signal transduction mechanisms and defense mechanisms in the intracellular bacteria of vertebrates (15.7%, Kruskal-Wallis test : p-value = 2.2*10-5 and 8.1%, 1.8*10-6, respectively; Correlation test : p-value = 5.5*10-2 and 3.1*10-3, respectively); the transport and metabolism of amino acid and coenzyme in bacteria from insects (15.6%, Kruskal-Wallis test: p-value = 1.0*10-2 and 13.3%, ANOVA test : p-value = 1.1*10-7, respectively), the transport and metabolism of coenzymes in bacteria from soils (9.7%, ANOVA test: p-value = 1.1*10-7) (Fig. 2). Although they had high proportions of specific genes, ubiquitous bacteria, bacteria from water and extracellular bacteria from vertebrates had no overrepresented functional category compared to the others (Fig. 2).

Fig. 2
figure 2

Functional distribution of the specific genes. a Proportions of each COG categories in the bacteria from different habitats. b The repartition of the four functional categories in each studied habitat. The red stars indicate the overrepresented functions in a habitat, compared to the other habitats, we also indicated the percent of specific genes supporting these functions

Horizontal transfers

Phylogenetic analyses conducted to infer the evolutionary origin of all the proteins other than the common genes indicated that 12,885 proteins (6.3% of all the proteins) were acquired via horizontal transfer (Fig. 3). As transfer events are not necessarily confined to individual genes but they may concern a cluster of genes, we considered neighboring genes with the same horizontal transfer history to reflect only a single event of HT. We counted a total of 10,918 HT events among studied bacteria. The incidence with which the HT events occurred was as high as 170.6 +/− 116.4, yet this was highly variable among bacteria ranging from 0 transfers in the genome of Blattabacterium and Borrelia spp. to 803 sequences transferred in the genome of Chthoniobacter (Fig. 3 and Additional file 6). The count of HT events was not correlated with the genome size (Wilcoxon, p = 0.042) (Additional file 7). The transferred fragments length ranged from 859 bp for C. tepidum to 1492 pb for S. acidophila, with a mean length of 1124+/− 274 bp for all fragments. The size of the HT was 1022 +/− 383 bp on average, and was similar in the different habitats. Bacteria from soil, from insects and ubiquitous bacteria presented the highest average of HT (524.0, 343.7 and 269.3 transfer events per species, respectively), compared to bacteria living in water, extracellular bacteria from vertebrates, bacteria from amoeba and the intracellular bacteria of vertebrates (183.7, 126.7, 113.5 and 17.6 HT per species, respectively). The statistical comparison of the bacterial group from different habitats allowed us to define three classes, based on percentages of sequences due to transfer (Kruskal-Wallis test : p-value = 2.3*10-2): Bacteria of soils (8.4%) and insects (7.3%) and ubiquitous bacteria (6.6%) defined the first class. Bacteria from soil-water (5.0%), water (4.9%), amoeba (4.8%), and extracellular bacteria of vertebrates (4.6%) presented similar proportions, and formed the second class. Intracellular bacteria of vertebrates were grouped in the third class, with 1.9% of HT events (Additional files 6).

Fig. 3
figure 3

Functional description of the detected horizontally transferred genes in each studied bacteria. The habitats of species and their ancestors are indicated by the color of the branches (red : ubiquitous, cyan : water, green : insects, yellow: amoeba, pink: intracellular vertebrates, purple: extracellular vertebrates, blue: water and soils, brown: soils), the black dotted branches indicate an unknown habitat. The compartmentalized bacteria are indicated in bold. The first column contains the number assigned to each species in Fig. 5, the second presents the genome sizes of bacteria, the third, the quantity of genes transferred in genomes, and the fourth, the quantity of HT detected. The bar graphic represents the distribution of the four functional categories of COGs (Information storage and processing, Metabolism, Cellular process and signaling, Poorly characterized), which contain the 18 sub-categories studied

Most of the horizontal exchanges were realized with bacteria (94.9%) and very few with Archaea (2.4%), Eukaryota (2.5%) and viruses (0.2%). Bacteria from amoeba and ubiquitous bacteria showed a significantly higher quantity of HT instances realized with eukaryotes, compared to bacteria from other habitats (9.8 and 4.2%, respectively) (Kruskal test : p-value = 3.3*10-3). The proportion of exchanges with Archaea was significantly higher in bacteria from water-soils and from water (3.4 and 3.4%, respectively) compared to bacteria from other habitats (Kruskal test: p-value = 3.1*10-2). The most common bacterial partners identified were Proteobacteria (42%), Firmicutes (23%) and Cyanobacteria or Actinobacteria (6%) (Fig. 4). For several transfer partners, we identified significant differences among the 64 bacteria studied, according to their habitat: bacteria from intracellular and extracellular vertebrates were both characterized by their preference for the Firmicutes (21.4 and 41.1%, respectively) as transfer partners (Kruskal test : p-value = 3.6*10-3), and their significantly lower proportion of transfers with Actinobacteria (ANOVA test : p-value = 3.8*10-2), compared to bacteria from other habitats. Extracellular bacteria from vertebrates also presented a significantly higher proportion of transfers with Fusobacteria (3.0%) compared to bacteria from other habitats (Kruskal test: p-value = 1.6*10-2), whereas the bacteria living in soil, or soil and water, exchanged significantly more with Acidobacteria (4.1 and 3.9% respectively) (Kruskal test: p-value = 5.3*10-4) (Fig. 4). The Principal components analysis (PCA) of data recovered for HT (HT proportions and partners) showed a relationship between bacterial habitats, the quantity of HT events, and its proportion in the genome and the partner transfers (Correlation test: p-value = 2.4*10-7). Hierarchical clustering analysis allowed the identification of two major clusters: the environmental bacteria (soils, water, soils-water and Ubiquitous bacteria) and bacteria from amoeba were in a first cluster, and the intracellular and extracellular bacteria of vertebrates were in the second cluster (Fig. 5). A clustering according to phylogenetic relationships among species can also be identified, but was less significant (Correlation test: p-value = 3.5*10-4) than clustering by habitat.

Fig. 4
figure 4

Preferences for horizontal transfer partners among habitats. The colored point corresponds to the bacteria from the different habitats (red : ubiquitous, cyan : water, green : insects, yellow: amoeba, pink: intracellular vertebrates, purple: extracellular vertebrates, blue: water and soils, brown: soils). The traits are colored according to the habitats of studied bacteria and their thickness is proportional to the amount of genes exchanged

Fig. 5
figure 5

Results for the principal component analysis and hierarchical clustering. They were realized with the variables: HT quantities and proportions in the studied genomes and their partners. The individuals represented are the 60 bacteria where HT were identified, the colors of points and numbers indicate their habitat. Two clusters were defined (transparent color squares) according the bacterial habitats (Chi2 test = 2.4*10-7): 91.3% of bacteria from vertebrates are gathered in the cluster 1 (blue square) and 92.0% of environmental bacteria belong to the cluster 2 (red square). Axis 1 contains 33.9% of the information and mainly represents the transfer partners (Variable represented: Cyanobacteria, Firmicutes, Fusobacteria and Alphaproteobacteria (>40%), Actinobacteria, Acidobacteria and Epsilonbacteria (>20%)). The axis 2 contains 21.0% of total information and represents mainly the HT quantities and proportions (Variables represented: Sequences quantities and proportions (>70%), Actinobacteria and Gammaproteobacteria (>15%)). However, 46.1% of the information is missing, concerning mainly the Gammaproteobacteria and Acidobacteria

Horizontally transferred functions and phenotype

When analyzing the functions of the transferred sequences, we found that the general function distribution in the HT for the different habitats was not similar to that of the whole genomes which suggests that HT was not due to chance. Genes involved in cell processes and signaling (33 to 50%) seemed to be significantly more subject to HT, whereas genes dedicated to information storage (12 to 17%) were less subject to HT (t-test between whole genomes and transferred genes : p-value = 4.6*10-2 and 8.3*10-4) (Fig. 3). Moreover, there were significant differences among the habitats. Biological functions of transferred sequences were biased to three categories according to bacterial habitat: the signal transduction mechanism function in ubiquitous bacteria and in bacteria from soils (20.2 and 17.9%, respectively; ANOVA test: p-value = 2.3*10-4) the transport and metabolism of amino acid in bacteria from amoeba and lipids in ubiquitous bacteria (16%, Kruskal-Wallis test and correlation test: p-value = 7.5*10-2 and 2.1*10-4; 10.5%, ANOVA test: p-value = 6.9*10-3, respectively) and the defense mechanism in bacteria from extracellular vertebrates (4.7%, Kruskal-Wallis: p-value = 3.5*10-2).

To test the impact of compartmentalization on HT, we compared the 11 compartmentalized with 9 non-compartmentalized bacteria in water, where the sample size allowed us to obtain statistically significant results (Fig. 3). HT proportions and preferences for partners were identical between the two groups of bacteria. However, we detected two functions that were overrepresented among the HT events of compartmentalized bacteria, compared to non-compartmentalized, including cell wall/membrane/envelope biogenesis (5.5% of transferred genes, ANOVA test: p-value = 5.05*10-2) and lipid biosynthesis (2.4% of transferred genes, ANOVA test : p-value = 3.7*10-2) (Fig. 6). Genes encoding for these two functions were found to be horizontally transferred in compartmentalized bacteria of the different habitats, including 25 genes implicated in cell wall/membrane/envelope biogenesis (of which 12 are present in at least, one compartmentalized bacteria from each phylum) and 13 genes implicated in lipid biosynthesis (of which 3 were present in at least one compartmentalized bacteria from each phylum). Moreover, five of the genes implicated in the two functions of interest were found to be horizontally transferred in at least one compartmentalized bacteria from each habitat. These genes encoded for two different glycosyl transferases (Gene ID: 1791906 and 1796578), a carboxy-terminal processing protease (Gene ID: 1796308), an ACP reductase (Gene ID: 1793322), and a Resistance-Nodulation-Cell Division (RND) efflux membrane fusion protein (Gene ID: 1796098).

Fig. 6
figure 6

Overrepresented functions among genes transferred in compartmentalized bacteria from water. Compartmentalized bacteria are represented on the left of figure, non compartmentalized bacteria are represented on the right. Color arrows on the sides indicate the functions with similar proportions in the two groups of bacteria, these proportions are indicated close to the arrows. Colored arrows, in the center, show the functions that are over-represented in compartmentalized bacteria, with their complete name and the p-value of comparison test

Discussion

The comparative analysis of 33 genomes from PVC species from four different phyla showed the influence of the living environment and compartmentalization on the genome composition of PVC bacteria. The common genes were genes encoding for transcription, signal transduction mechanisms, energy production and membrane biogenesis. Conversely, shared and specific genes encode for different functions in relation to the lifestyle of the corresponding species. Evidence for a random horizontal transfer of DNA sequences has been given using a phylogenomic approach. Genes implicated in cell wall/membrane/envelope biogenesis, and those involved in lipid metabolism, were found to be over-represented among the transferred genes of compartmentalized bacteria from different habitats, according to a convergent evolutionary selection.

Our findings replicate observations from previous studies which demonstrated the role played by shared genomes in environmental adaptation [68]. Nevertheless, our approach, by examining as many as 8 different habitat conditions, offers a large advantage over other genomic studies, and increases the reliability of our results. The low proportion of specific genes that have been detected in bacteria from insects, soils and soil-water milieu is rather due to the higher number and more distant phylogenetic relationships among the species studied [69, 70] compared to the other habitats. Intracellular bacteria from vertebrates showed a low proportion (1.9%) of horizontally transferred sequences compared to the other bacteria. This result is probably related to the physical isolation of intracellular bacteria, which prevents opportunities for HT [71, 72]. This agrees with previous studies showing that the predominant evolutionary process in intracellular bacteria is genome reduction, leading to smaller genome sizes [73, 74]. Intracellular bacteria in amoeba with 4.8% of HT are the exception [75, 76], since amoeba can phagocyte several bacteria at once, giving a particular field for potential genetic exchange and a training ground for the emergence of parasitism [75].

Likewise, results obtained for partners of transfers analysis were in agreement with previous results concerning transfers between PVC bacteria and Proteobacteria [43] or Spirochaetes and Firmicutes [12, 13]. Indeed, HT occurred preferentially between bacteria from the same habitats, as had already been assumed. Firmicutes are one of the two major phyla present in the gut microbiome [77], and this is the main partner of our bacteria from vertebrates. In the same way, Acidobacteria are mainly detected in soil [78] and they are overrepresented as HT partners of bacteria from soils, compared to bacteria from other habitats. The tendency of bacteria from Amoeba to exchange more with Eukaryotes, especially plants, is probably due to their ancestral habitats. Indeed, ancestral Chlamydiae are known to have lived in and exchanged genes with the Archaeplastides [79, 80]. Thus, we can support the hypothesis that part of the HT detected was acquired by the interaction between the ancestors of the Chlamydiae and the plants, followed by the loss in the majority of bacteria. It is worth noting that like previous studies for HT detections, it is difficult to distinguish between ancient and recent HT events; yet HT partners are the witnesses of modern and ancestral habitats of the bacteria studied, and our HT analysis helps infer the ancestral habitat of these bacteria.

Beyond the complexity hypothesis that claims that genes involved in transcription and translation are less prone to transfer than metabolic genes, our findings showed that horizontal transfers can affect any function. Thus, HT do not only concern genes encoding for metabolic mechanisms and other functions that enhance pathogenicity, like genes for virulence and antimicrobial activity [1, 15, 76]; genes involved in transcription and translation, in cell surface and DNA binding, and genes essential for defense can likely be transferred as well [46, 8183]. Positive selection might be contributing to the over-representation of some functions in the category of transferable genes [84, 85]. Indeed, horizontally acquired genes that have a useful function are maintained as it follows a strategy of colonization and adaptation to the environment. Our findings confirm previous results showing that HT particularly affects the genes involved in lipid metabolism, signal transduction and membrane transport in PVC bacteria, and genes specific to outer membrane (such as O-antigen polymerase and outer membrane efflux protein) in some Planctomycetes [43, 86, 87]. Since, the intracytoplasmic membrane of compartmentalized bacteria is a lipid bilayer, we can assume that the over-representation of the two functions in the genes transferred could be related to the cell plan of the bacteria. These genes may be essential for the maintenance of the supplementary intracytoplasmic membrane. Knowing that the quantity of HT events was found to be similar between compartmentalized and non compartmentalized bacteria, these results revealed the possible impact of the cell plan on the transfers’ positive selection. This selection that seems to be dependent on the function, and induces the recurrent maintenance of some transferred genes involved in the formation of compartments in bacteria from different habitats. It is noteworthy that genes implicated in lipid metabolism and membrane biosynthesis were not over-represented in the non-transferred part of the genome of compartmentalized bacteria, compared to the other bacteria; therefore, the selection seems to concern only the transferred genes.

One limitation of our comparative genomic approach is that the number of genomes studied leads to a small sample size in each environmental category, which hinders the realization of the statistical tests for certain categories. Moreover, our dataset was comprised of seven phyla, with only few representatives of soil bacteria, four for bacteria living in amoeba, and three extracellular bacteria from insects or ubiquitous, while some environmental categories contain only bacteria from just one phylum. Although the sample size was minimal, the results obtained were statistically usable and showed significant differences among phylogenetically close bacteria in relation with their habitat. Given the increased number of sequenced genomes, it will be interesting to characterize HT events in compartmentalized bacteria for diverse phyla, in order to elucidate the role of physical barriers in horizontal transfers.

Conclusions

The genomic study of bacteria allowed to better understand the influence of the different constraints acting on genomes evolution in bacteria, especially the impact of the habitat and the special cell plan, in PVC super-phylum. The habitat influences the flux of horizontal transfer and determines the partners for genetic exchanges. The presence of an intracytoplasmic membrane in some PVC bacteria doesn’t seem to limit the HT but rather, induces a selection of transferred genes, according to their functions.

References

  1. Ochman H, Lawrence JG, Groisman EA. Lateral gene transfer and the nature of bacterial innovation. Nature. 2000;405(6784):299–304.

    Article  CAS  PubMed  Google Scholar 

  2. Le PT, Ramulu HG, Guijarro L, Paganini J, Gouret P, Chabrol O, et al. An automated approach for the identification of horizontal gene transfers from complete genomes reveals the rhizome of Rickettsiales. BMC Evol Biol. 2012;12(1):243.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Thomas CM, Nielsen KM. Mechanisms of, and barriers to, horizontal gene transfer between bacteria. Nat Rev Microbiol. 2005;3(9):711–21.

    Article  CAS  PubMed  Google Scholar 

  4. Ghigo JM. Natural conjugative plasmids induce bacterial biofilm development. Nature. 2001;412:442–5.

    Article  CAS  PubMed  Google Scholar 

  5. Reisner A, Höller BM, Molin S, Zechner EL. Synergistic effects in mixed Escherichia coli biofilms: conjugative plasmid transfer drives biofilm expansion. J Bacteriol. 2006;188:3582–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Aminov RI. Horizontal gene exchange in environmental microbiota. Front Microbiol. 2011;2(158):10–3389.

    Google Scholar 

  7. Smillie CS, Smith MB, Friedman J, Cordero OX, David LA, Alm EJ. Ecology drives a global network of gene exchange connecting the human microbiome. Nature. 2011;480(7376):241–4.

    Article  CAS  PubMed  Google Scholar 

  8. Ilves H, Horak R, Kivisaar M. Involvement of sigma (S) in starvation-induced transposition of Pseudomonas putida transposon Tn4652. J Bacteriol. 2001;183:5445–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Dorer MS, Fero J, Salama NR. DNA damage triggers genetic exchange in Helicobacter pylori. PLoS Pathog. 2010;6(7):e1001026.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Stanton TB, Humphrey SB, Sharma VK, Zuerner RL. Collateral effects of antibiotics: carbadox and metronidazole induce VSH-1 and facilitate gene transfer among Brachyspira hyodysenteriae strains. Appl Environ Microbiol. 2014;74:2950–6.

    Article  Google Scholar 

  11. Hacker J, Carniel E. Ecological fitness, genomic islands and bacterial pathogenicity : A Darwinian view of the evolution of microbes. EMBO Rep. 2011;21:376–81.

    Google Scholar 

  12. Bellgard MI, Wanchanthuek P, La T, Ryan K, Moolhuijzen P, Albertyn Z. Genome sequence of the pathogenic intestinal spirochete brachyspira hyodysenteriae reveals adaptations to its lifestyle in the porcine large intestine. PLoS One. 2009;4(3):e4641.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Viswanathan VK. Spirochaeta and their twisted ways. Gut Microbes. 2012;3(5):399–400.

    Article  PubMed  Google Scholar 

  14. Craigie R, Gellert M, Lambowitz AM. Mobile DNA. In: Craig NL, editor. American society for microbiology. 1989.

    Google Scholar 

  15. Baquero F, Alvarez‐Ortega C, Martinez JL. Ecology and evolution of antibiotic resistance. Environ Microbiol Rep. 2009;1(6):469–76.

    Article  CAS  PubMed  Google Scholar 

  16. Gemski P, Lazere JR, Casey T, Wohlhieter JA. Presence of virulence-associated plasmid in Yersinia pseudotuberculosis. Infect Immun. 1980;28:1044–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Groisman EA, Ochman H. Pathogenicity islands: bacterial evolution in quantum leaps. Cell. 1996;87:791–4.

    Article  CAS  PubMed  Google Scholar 

  18. Hacker J, Blum-Oehler G, Muhldorfer I, Tschape H. Pathogenicity islands of virulent bacteria: structure, function and impact on microbial evolution. Mol Microbiol. 1997;23:1089–97.

    Article  CAS  PubMed  Google Scholar 

  19. Marraffini LA, Sontheimer EJ. CRISPR interference: RNA-directed adaptive immunity in bacteria and archaea. Nat Rev Genet. 2010;11(3):181–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Bikard D, Hatoum-Aslan A, Mucida D, Marraffini LA. CRISPR interference can prevent natural transformation and virulence acquisition during in vivo bacterial infection. Cell Host Microbe. 2012;12(2):177–86.

    Article  CAS  PubMed  Google Scholar 

  21. González-Candelas F, Francino MP. Barriers to Horizontal Gene Transfer: Fuzzy and Evolvable Boundaries. In: Pilar F editor. Horizontal Gene Transfer in Microorganisms. Caister: Academic Press. 2012. p. 47.

  22. Besemer K, Singer G, Quince C, Bertuzzo E, Sloan W, Battin TJ. Headwaters are critical reservoirs of microbial diversity for fluvial networks. Proc R Soc Lond B Biol Sci. 2013;280(1771):20131760.

    Article  Google Scholar 

  23. Maccario L, Vogel TM, Larose C. Potential drivers of microbial community structure and function in Arctic spring snow. Front Microbiol. 2014;5:413.

    Article  PubMed  PubMed Central  Google Scholar 

  24. Mendes LW, Tsai SM, Navarrete AA, de Hollander M, van Veen JA, Kuramae EE. Soil-borne microbiome: linking diversity to function. Microb Ecol. 2015;70(1):1–11.

    Article  Google Scholar 

  25. Caro-Quintero A, Konstantinidis KT. Inter-phylum HGT has shaped the metabolism of many mesophilic and anaerobic bacteria. ISME J. 2015;9(4):958–67.

    Article  CAS  PubMed  Google Scholar 

  26. Tyson GW, Chapman J, Hugenholtz P, Allen EE, Ram RJ, Richardson PM, et al. Community structure and metabolism through reconstruction of microbial genomes from the environment. Nature. 2004;428(6978):37–43.

    Article  CAS  PubMed  Google Scholar 

  27. Lauro FM, McDougald D, Thomas T, Williams TJ, Egan S, Rice S, et al. The genomic basis of trophic strategy in marine bacteria. Proc Natl Acad Sci. 2009;106(37):15527–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Cho JC, Vergin KL, Morris RM, Giovannoni SJ. Lentisphaera araneosa gen. nov., sp. nov, a transparent exopolymer producing marine bacterium, and the description of a novel bacterial phylum, Lentisphaerae. Environ Microbiol. 2004;6(6):611–21.

    Article  CAS  PubMed  Google Scholar 

  29. Wagner M, Horn M. The Planctomycetes, Verrucomicrobia, Chlamydiae and sister phyla comprise a superphylum with biotechnological and medical relevance. Curr Opin Biotechnol. 2006;17(3):241–9.

    Article  CAS  PubMed  Google Scholar 

  30. Siegl A, Kamke J, Hochmuth T, Piel J, Richter M, Liang C, et al. Single-cell genomics reveals the lifestyle of Poribacteria, a candidate phylum symbiotically associated with marine sponges. ISME J. 2011;5(1):61–70.

    Article  PubMed  Google Scholar 

  31. Gupta RS, Bhandari V, Naushad HS. Molecular signatures for the PVC clade (Planctomycetes, Verrucomicrobia, Chlamydiae, and Lentisphaerae) of bacteria provide insights into their evolutionary relationships. Front Microbiol. 2012;3:327.

    PubMed  PubMed Central  Google Scholar 

  32. Lindsay MR, Webb RI, Fuerst JA. Pirellulosomes: a new type of membrane-bounded cell compartment in planctomycete bacteria of the genus Pirellula. Microbiology. 1997;143(3):739–48.

    Article  CAS  Google Scholar 

  33. Lindsay MR, Webb RI, Strous M, Jetten MS, Butler MK, Forde RJ, Fuerst JA. Cell compartmentalisation in planctomycetes: novel types of structural organisation for the bacterial cell. Arch Microbiol. 2001;175(6):413–29.

    Article  CAS  PubMed  Google Scholar 

  34. Fuerst JA, Sagulenko E. Beyond the bacterium: planctomycetes challenge our concepts of microbial structure and function. Nat Rev Microbiol. 2011;9:403–13.

    Article  CAS  PubMed  Google Scholar 

  35. Lee KC, Webb RI, Janssen PH, Sangwan P, Romeo T, Staley JT, Fuerst JA. Phylum Verrucomicrobia representatives share a compartmentalized cell plan with members of bacterial phylum Planctomycetes. BMC Microbiol. 2009;9(1):5.

    Article  PubMed  PubMed Central  Google Scholar 

  36. Fieseler L, Horn M, Wagner M, Hentschel U. Discovery of the novel candidate phylum “Poribacteria” in marine sponges. Appl Environ Microbiol. 2004;70(6):3724–32.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Fuerst JA, Webb RI. Membrane-bounded nucleoid in the eubacterium Gemmatata obscuriglobus. Proc Natl Acad Sci. 1991;88(18):8184–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Lage OM, Bondoso J, Lobo-da-Cunha A. Insights into the ultrastructural morphology of novel Planctomycetes. Antonie Van Leeuwenhoek. 2013;104(4):467–76.

    Article  PubMed  Google Scholar 

  39. Santarella-Mellwig R, Franke J, Jaedicke A, Gorjanacz M, Bauer U, Budd A, Devos DP. The compartmentalized bacteria of the planctomycetes-verrucomicrobia-chlamydiae superphylum have membrane coat-like proteins. PLoS Biol. 2010;8(1):e1000281.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Santarella-Mellwig R, Pruggnaller S, Roos N, Mattaj IW, Devos DP. Three-dimensional reconstruction of bacteria with a complex endomembrane system. PLoS Biol. 2013;11(5):e1001565.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. NCBI - proteins.https://www.ncbi.nlm.nih.gov/protein. Accessed January and March 2013.

  42. NCBI - genomes.http://mirrors.vbi.vt.edu/mirrors/ftp.ncbi.nih.gov/genomes/. Accessed January and March 2013.

  43. Kamneva OK, Knight SJ, Liberles DA, Ward NL. Analysis of genome content evolution in PVC bacterial super-phylum: assessment of candidate genes associated with cellular organization and lifestyle. Genome Biol Evol. 2012;4(12):1375–90.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Tamura K, Peterson D, Peterson N, Stecher G, Nei M, Kumar S. MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol Biol Evol. 2011;28:2731–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004;32(5):1792–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Merhej V, Notredame C, Royer-Carenzi M, Pontarotti P, Raoult D. The rhizome of life: the sympatric Rickettsia felis paradigm demonstrates the random transfer of DNA sequences. Mol Biol Evol. 2011;28(11):3213–23.

    Article  CAS  PubMed  Google Scholar 

  47. Georgiades K. Genomics of epidemic pathogens. Clin Microbiol Infect. 2013;18(3):213–7.

    Article  Google Scholar 

  48. JGI - genomes. http://genome.jgi.doe.gov/. Accessed between August and December 2014.

  49. GOLD database. https://gold.jgi-psf.org/index. Accessed between August and December 2014.

  50. List of Prokaryotic names with Standing in Nomenclature - Bacterio.net. http://www.bacterio.net/index.html. Accessed between August and December 2014.

  51. Grigoriev IV, Nordberg H, Shabalov I, Aerts A, Cantor M, Goodstein D, et al. The genome portal of the department of energy joint genome institute. Nucleic Acids Res. 2012;D1(40):D26–32.

    Article  Google Scholar 

  52. Nordberg H, Cantor M, Dusheyko S, Hua S, Poliakov A, Shabalov I, et al. The genome portal of the Department of Energy Joint Genome Institute: 2014 updates. Nucleic Acids Res. 2014;42(D1):D26–31.

    Article  CAS  PubMed  Google Scholar 

  53. Reddy TBK, Thomas AD, Stamatis D, Bertsch J, Isbandi M, Jansson J, et al. The Genomes OnLine Database (GOLD) v. 5: a metadata management system based on a four level (meta) genome project classification. Nucleic Acids Res. 2014;43(D1):D1099–D110.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Pinos S, Pontarotti P, Raoult D, Baudoin JP, Pagnier I. Compartmentalization in PVC super-phylum: evolution and impact. Biol Direct. 2016;11(1):38.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Fischer S, Brunk BP, Chen F, Gao X, Harb OS, Iodice JB, et al. Using OrthoMCL to assign proteins to OrthoMCL‐DB groups or to cluster proteomes into New ortholog groups. Curr Protoc Bioinformatics. 2011;35(Suppl 6.12):1–19.

    Google Scholar 

  56. Tatusov RL, Koonin EV, Lipman DJ. A genomic perspective on protein families. Science. 1997;278(5338):631–7.

    Article  CAS  PubMed  Google Scholar 

  57. Wu S, Zhu Z, Fu L, Niu B, Li W. WebMGA: a customizable web server for fast metagenomic sequence analysis. BMC Genomics. 2011;12(1):444.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Galperin MY, Makarova KS, Wolf YI, Koonin EV. Expanded microbial genome coverage and improved protein family annotation in the COG database. Nucleic Acids Res. 2015;43(D1):D261–9.

    Article  PubMed  Google Scholar 

  59. Mitchell A, Chang HY, Daugherty L, Fraser M, Hunter S, Lopez R, et al. The InterPro protein families database: the classification resource after 15 years. Nucleic Acids Res. 2015;43(D1):D213–21.

    Article  PubMed  Google Scholar 

  60. Gouret P, Thompson JD, Pontarotti P. PhyloPattern: regular expressions to identify complex patterns in phylogenetic trees. BMC Bioinf. 2009;10(1):298.

    Article  Google Scholar 

  61. Sankoff D, Morel C, Cedergren RJ. Evolution of 5S RNA and the non-randomness of base replacement. Nat New Biol. 1973;245:232–4.

    Article  CAS  PubMed  Google Scholar 

  62. Levene H. Robust tests for equality of variance. In: Olkin I, Ghurye SG, Hoeffeling W, Madow WG, Mann HB, editors. Contributions to Probability and Statistics: Essays in Honor of Harold Hotelling. Stanford: University Press; 1960. p. 278–292.

  63. Kruskal WH, Wallis WA. Use of ranks in one-criterion variance analysis. J Am Stat Assoc. 1952;47(260):583–621.

    Article  Google Scholar 

  64. Douglas CE, Michael FA. On distribution-free multiple comparisons in the one-way analysis of variance. Commun Stat Theory Methods. 1991;20(1):127–39.

    Article  Google Scholar 

  65. Tukey JW. Comparing individual means in the analysis of variance. Biometrics. 1949;5:99–114.

    Article  CAS  PubMed  Google Scholar 

  66. Pagel M. Detecting correlated evolution on phylogenies: a general method for the comparative analysis of discrete characters. Proc R Soc London. 1994;B 255:37–45.

    Article  Google Scholar 

  67. Spearman C. The proof and measurement of association between two things. Am J Psychol. 1904;15:72–101.

    Article  Google Scholar 

  68. Tettelin H, Riley D, Cattuto C, Medini D. Comparative genomics: the bacterial pan-genome. Curr Opin Microbiol. 2008;11:472–7.

    Article  CAS  PubMed  Google Scholar 

  69. Daubin V, Gouy M, Perriere G. A phylogenomic approach to bacterial phylogeny: evidence of a core of genes sharing a common history. Genome Res. 2002;12(7):1080–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Charlebois RL, Clarke GP, Beiko RG, Jean AS. Characterization of species-specific genes using a flexible, web-based querying system. FEMS Microbiol Lett. 2003;225(2):213–20.

    Article  CAS  PubMed  Google Scholar 

  71. Ley RE, Peterson DA, Gordon JI. Ecological and evolutionary forces shaping microbial diversity in the human intestine. Cell. 2006;124:837–48.

    Article  CAS  PubMed  Google Scholar 

  72. Hirt RP, Alsmark C, Embley TM. Lateral gene transfers and the origins of the eukaryote proteome: a view from microbial parasites. Curr Opin Microbiol. 2015;23:155–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Merhej V, Royer-Carenzi M, Pontarotti P, Raoult D. Massive comparative genomic analysis reveals convergent evolution of specialized bacteria. Biol Direct. 2009;4(1):13.

    Article  PubMed  PubMed Central  Google Scholar 

  74. Wolf YI, Koonin EV. Genome reduction as the dominant mode of evolution. Bioessays. 2013;35(9):829–37.

    Article  PubMed  PubMed Central  Google Scholar 

  75. Molmeret M, Horn M, Wagner M, Santic M, Kwaik YA. Amoebae as training grounds for intracellular bacterial pathogens. Appl Environ Microbiol. 2005;71(1):20–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Moliner C, Fournier PE, Raoult D. Genome analysis of microorganisms living in amoebae reveals a melting pot of evolution. FEMS Microbiol Rev. 2010;34(3):281–94.

    Article  CAS  PubMed  Google Scholar 

  77. Qin J, Li R, Raes J, Arumugam M, Burgdorf KS, Manichanh C, et al. A human gut microbial gene catalogue established by metagenomic sequencing. Nature. 2010;464(7285):59–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Quaiser A, Ochsenreiter T, Lanz C, Schuster SC, Treusch AH, Eck J, Schleper C. Acidobacteria form a coherent but highly diverse group within the bacterial domain: evidence from environmental genomics. Mol Microbiol. 2006;50(2):563–75.

    Article  Google Scholar 

  79. Subtil A, Collingro A, Horn M. Tracing the primordial Chlamydia: extinct parasites of plants? Trends Plant Sci. 2014;19(1):36–43.

    Article  CAS  PubMed  Google Scholar 

  80. Moustafa A, Reyes-Prieto A, Bhattacharya D. Chlamydia has contributed at least 55 genes to Plantae with predominantly plastid functions. PLoS One. 2008;3(5):e2205.

    Article  PubMed  PubMed Central  Google Scholar 

  81. Tooming-Klunderud A, Sogge H, Rounge TB, Nederbragt AJ, Lagesen K, Glöckner G, et al. From green to red: Horizontal gene transfer of the phycoerythrin gene cluster between Planktothrix strains. Appl Environ Microbiol. 2013;79(21):6803–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Nakamura Y, Itoh T, Matsuda H, Gojobori T. Biased biological functions of horizontally transferred genes in prokaryotic genomes. Nat Genet. 2004;36(7):760–6.

    Article  CAS  PubMed  Google Scholar 

  83. Mongodin EF, Nelson KE, Daugherty S, DeBoy RT, Wister J, Khouri H, et al. The genome of Salinibacter ruber: Convergence and gene exchange among hyperhalophilic bacteria and archaea. PNAS. 2005;102(50):18147–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Gogarten JP, Doolittle WF, Lawrence J. G Prokaryotic evolution in light of gene transfer. Mol Biol Evol. 2002;19:2226–38.

    Article  CAS  PubMed  Google Scholar 

  85. Pál C, Papp B, Lercher MJ. Horizontal gene transfer depends on gene content of the host. Bioinformatics. 2005;21 Suppl 2:ii222–3.

    PubMed  Google Scholar 

  86. Fuerst JA. The PVC superphylum: exceptions to the bacterial definition? Antonie Van Leeuwenhoek. 2013;104(4):451–66.

    Article  CAS  PubMed  Google Scholar 

  87. Paparoditis P, Västermark Å, Le AJ, Fuerst JA, Saier MH. Bioinformatic analyses of integral membrane transport proteins encoded within the genome of the planctomycetes species, Rhodopirellula baltica. Biochim Biophys Acta Biomembr. 2014;1838(1):193–215.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank Olivier Chabrol for assistance in computer programming during the elaboration of HT detection strategy and Manuela Royer Carenzi for her assistance in the statistical analyzes. We also thank the Xegen company for their assistance in HT detection by using of Phylopattern software. We thank TradOnline for English reviewing.

Funding

This work was supported by the Assistance Publique - Hopitaux de Marseille (Marseille Public University Hospital System). VM was supported by a Chairs of Excellence program from the Centre National de la Recherche Scientifique (CNRS). The funders had no role in study design, data collection and interpretation or the decision to submit the work.

Availability of data and materials

The dataset supporting the conclusions of this article is included within the article and its additional files.

Authors’ contributions

PS carried out the design of the study, the strategy elaboration and the collection of data, performed the statistical analysis of results, and drafted the manuscript. PP participated in strategy elaboration, data interpretation and revised the manuscript. DR conceived the study, participated in its design and coordination and revised the manuscript. VM participated in the coordination of the study, strategy elaboration and the interpretation of data, and also drafted the manuscript. All authors read and approved the final manuscript.

Competing interests

The authors declare that they have no competing interests.

Consent for publication

Not applicable.

Ethics approval and consent to participate

Not applicable.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vicky Merhej.

Additional files

Additional file 1:

Genomic features and habitats of bacteria studied. For each of the bacteria studied, we present the genomes identifiers, genomic features (proteins quantity, GC percent), the habitat (environment and lifestyle) and the cell plan. (CSV 6 kb)

Additional file 2:

Proteins used for phylogenetic reconstruction. For each of the 768 sequences used for alignments and phylogenies we retrieved the accession number (column 1), the name of marker (column 2) and the species (column 3). (CSV 37 kb)

Additional file 3:

Phylogeny of PVC bacteria and phylogeny of Bacteroidetes-Chlorobi-Spirochaetes. The phylogenies of studied bacteria were realized with Maximum likelihood method, within Mega 5 software, the bootstraps values are indicated at each node. The trees of PVC bacteria and Bacteroidetes-Spirochaetes-Chlorobi were rooted with Bacteroides xylanosolvens (Spirochaetes) and with Bastopirellula marina (Planctomycetes), respectively (the outgroup was removed for more clear representation). (TIF 2084 kb)

Additional file 4:

The two phylogenetic patterns used for HT detection. The presence of one of these two patterns in a species tree allows us to identify an HT, these patterns are illustrated with examples specific to our data. (TIF 898 kb)

Additional file 5:

Clustering of species according orthologous groups distribution. Clustering of studied species, according the pattern of presence/absence of genes in orthologous groups. The habitats of species are indicated by colors (red: ubiquitous, cyan: water, green: insects, yellow: amoeba, pink: intracellular vertebrates, purple: extracellular vertebrates, blue: water and soils, brown: soils) and phylum of species are in brackets. (TIF 1303 kb)

Additional file 6:

HT distribution in habitats. The different boxplots describe the distribution of the numbers (a) and the proportions (b) of HT in each habitat. (TIF 878 kb)

Additional file 7:

HT count according to genome size in studied bacteria. The trend curve (linear), its function (y) and the determination of coefficient (R2) are indicated on the plot. (TIF 606 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

Pinos, S., Pontarotti, P., Raoult, D. et al. Identification of constraints influencing the bacterial genomes evolution in the PVC super-phylum. BMC Evol Biol 17, 75 (2017). https://doi.org/10.1186/s12862-017-0921-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12862-017-0921-3

Keywords