Next Article in Journal
Development of a Generic PBK Model for Human Biomonitoring with an Application to Deoxynivalenol
Previous Article in Journal
Incidence of Aflatoxins and Ochratoxin A in Wheat and Corn from Albania
Previous Article in Special Issue
Realized Heritability, Risk Assessment, and Inheritance Pattern in Earias vittella (Lepidoptera: Noctuidae) Resistant to Dipel (Bacillus thuringiensis Kurstaki)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recombination in Bacterial Genomes: Evolutionary Trends

by
Anton E. Shikov
1,2,
Iuliia A. Savina
1,
Anton A. Nizhnikov
1,2 and
Kirill S. Antonets
1,2,*
1
Laboratory for Proteomics of Supra-Organismal Systems, All-Russia Research Institute for Agricultural Microbiology (ARRIAM), 196608 St. Petersburg, Russia
2
Faculty of Biology, St. Petersburg State University (SPbSU), 199034 St. Petersburg, Russia
*
Author to whom correspondence should be addressed.
Toxins 2023, 15(9), 568; https://doi.org/10.3390/toxins15090568
Submission received: 8 August 2023 / Revised: 2 September 2023 / Accepted: 7 September 2023 / Published: 12 September 2023

Abstract

:
Bacterial organisms have undergone homologous recombination (HR) and horizontal gene transfer (HGT) multiple times during their history. These processes could increase fitness to new environments, cause specialization, the emergence of new species, and changes in virulence. Therefore, comprehensive knowledge of the impact and intensity of genetic exchanges and the location of recombination hotspots on the genome is necessary for understanding the dynamics of adaptation to various conditions. To this end, we aimed to characterize the functional impact and genomic context of computationally detected recombination events by analyzing genomic studies of any bacterial species, for which events have been detected in the last 30 years. Genomic loci where the transfer of DNA was detected pertained to mobile genetic elements (MGEs) housing genes that code for proteins engaged in distinct cellular processes, such as secretion systems, toxins, infection effectors, biosynthesis enzymes, etc. We found that all inferences fall into three main lifestyle categories, namely, ecological diversification, pathogenesis, and symbiosis. The latter primarily exhibits ancestral events, thus, possibly indicating that adaptation appears to be governed by similar recombination-dependent mechanisms.
Key Contribution: The presented review summarizes complex recombination-driven evolutionary routes of bacterial genomes with concomitant functional ramifications enabling adaptation to the environment.

1. Introduction

Through shaping the genomic landscape, horizontal gene transfer (HGT) or lateral gene transfer (LGT) and homologous recombination (HR) both serve as principal evolutionary forces in bacteria. These mechanisms provide genetic plasticity and thereby ensure adaptation to ecological niches [1], regulate virulence [2] and increase fitness [3]. The effect of these events is not only constrained to bacteria, but also plays a vital role in orchestrating the evolution and adaptability of archaea, viruses, and even eukaryotes [4]. The first phenomenon mentioned implies the replacement of DNA sequences when affecting genomic loci with contiguous, highly homologous regions [5]. HGT, in its turn, could be roughly defined as the acquisition of genetic material from a donor to recipient bacterial cells, predominantly requiring micro-homology sufficient for the incorporation of exogenous DNA into bacterial chromosomes or plasmids [6,7]. Not only do HR and HGT entail speciation [8] and spark the origin of new strains [4], but also cause antibiotic resistance [9], enhanced virulence [10], and reduced efficiency of vaccines [11]. However, the outcomes of these processes, in some cases, are beneficial for industries insofar as they result in the modulation of symbiotic relationships with agriculturally important plants [12] or lead to the emergence of strains capable of metabolizing pollutants, thus exhibiting promising biotechnological potential [13].
In terms of the parts of the bacterial genome affected, HR was reported to exert an effect on core genes [14], i.e., those shared by the majority of isolates within a certain population, whereas the accessory component is commonly embedded into genomic regions via HGT [15]. Therefore, the former alters allelic diversity, and the latter modifies gene composition. The preliminary step preceding the import or insertion of loci is DNA acquisition. Foreign DNA could enter bacterial cells via three fundamental mechanisms, namely, transformation, transduction, and conjugation [16]. Transformation involves a direct influx of DNA from the environment and is observed in diverse pathogens, including the genera Neisseria, Helicobacter, Streptococcus, etc. [7]. Being a complex multi-stage procedure, it is comprised of multiple steps, such as the activation of competence in the early stages of bacterial growth, triggering a two-component system ComD/ComE, and the action of the type IV pilus apparatus, through which foreign DNA transfer is performed [7]. The import is accompanied by DNA processing into 6 kb fragments by surface endonuclease [7]. Transduction is a phage-mediated DNA transfer between cells, whereby the phage carrying genetic material can integrate into a host genome [7]. The embedded loci could represent virulence determinants, e.g., toxin-encoding genes [17]. Conjugation, first discovered in E. coli when characterizing self-replicating F-plasmids, is carried out through direct cell-to-cell contact [16]. This necessity for maintaining cell contacts does not allow the transmission of genetic material other than small genetic elements, e.g., plasmids. It may also regulate the transfer of integrative and conjugative elements (ICEs), which lack a self-replication system and are capable of copying only during conjugation [7,16].
Upon obtaining DNA, either homologous or site-specific recombination occurs, with the latter governing HGT. HR is accompanied by the formation and consequent resolution of the Holliday junction [18]. It requires the RecA protein with polymerase activity attached to single-stranded DNA stabilized with SSB (single-stranded DNA-binding protein) [18]. Two alternative protein complexes, RecBCD and RecFOR, participate in the downstream stages [7,19]. Noteworthy, bacterial genomes demonstrate a non-uniform distribution of recombination systems with the prevalence of RecFOR over RecBCD in some organisms, presumably indicating the redundancy of the molecular mechanisms governing recombination [19]. Alternatively, obligate endosymbionts lack the RecA protein, while in their genomes, RecA-independent recombination has been reported [19,20,21], which can probably be explained by tandem repeats inducing genetic exchange [22]. HGT, understood as the incorporation of mobile genetic elements (MGEs) including plasmids, prophages, ICEs, and pathogenicity islands, is mainly mediated by site-specific recombination [7]. The respective integration is carried out by tyrosine recombinases, also known as integrases, and serine recombinases called resolvases [18]. The first family comprises XerC and XerD proteins, which separate chromosomal dimers during bacterial replication occurring at the dif site [18], whereas serine recombinases operate by making four breaks in double-stranded DNA molecules with their subsequent ligation [7].
HR and HGT are tightly interconnected, given that horizontally acquired genes are often flanked with genomic regions with an excessive HR frequency. This possibly serves to regulate genome size through excising obtained genes [23,24]. Additionally, transmitted MGEs could further engage in HR, as shown for prophages in cases of co-infection [25]. Moreover, pathogenic islands, insertion sequences, and other imported loci are characterized by HR signals predicted through bioinformatics tools [9,26,27,28]. On this account, in the current review, we would not distinguish between site-specific recombination-driven HGT and HR, using an umbrella term recombination instead.
Comparative studies have examined the intensity of HGT and HR to be unequal within bacterial populations that belong to different groups according to the ecological niches they occupy. As an example, obligate pathogens and symbionts are characterized by their lower HR rates than free-living organisms, commensals, and opportunistic pathogens [29,30]. Similar results held when considering horizontally transferred genes being presented by infection effectors and antibiotic resistance factors, thus, serving as modulators of bacterial invasion and promoting adaptation to certain hosts [3]. Taking into consideration the above-mentioned information, an in-depth understanding of the functional effect that recombination exerts on bacterial populations is needed concerning fundamental science and practical implications. In the review presented, we analyze aspects of recombination by discussing studies made in the last 30 years. The included studies comprised 91 bacterial species from different taxonomic groups. When choosing the articles, we focused on those in which recombination signals were found through computational predictions in whole genomes and/or individual loci, contiguous regions, and extrachromosomal elements. We summarize which parts of the genome are subjected to recombination and provide a scheme illustrating the biological roles of proteins encoded by the respective chimeric or acquired loci (Figure 1). We reveal that the vast majority of observations (Table S1) fall into three categories related to three extremely dynamic processes: the establishment and development of (i) symbiotic or (ii) pathogenic relationships and (iii) ecological diversification caused by alterations in the environment.

2. Functional Impact of Recombination

2.1. Ecological Adaptation

Here, ecological adaptation and diversification mean the successful transfer to different ecotopes, or even new niches accompanied by genomic diversification leading to specification. Recombination tends to occur when bacterial populations occupy geographically isolated regions, as shown for Alteromonas macleodii at different depths [1], Bacillus cereus/B. thuringiensis isolates in Norway [31], Arthrobacter sp. in glaciers [32], the freshwater bacterium Polynucleobacter asymbioticus [33], and Thermotoga maritima subpopulations in oil reservoirs [34]. Competitiveness is mainly maintained by either obtaining metabolic pathways for catabolizing new energy sources or maintaining the diversity of synthesized toxic compounds. For example, recombination allowed Pseudomonas stutzeri [13] and Arthrobacter nicotinovorans [35] to metabolize naphthalene and nicotine, respectively. Another recombinational scenario leads to strain diversity based on the synthesized toxins, so-called chemotypes, namely, nonribosomal toxic peptides of Planktothrix sp. [36] and Microcystis sp. [37] or the diverse antimicrobial secondary metabolites of Streptomyces sp. [38].

2.2. Symbiotic Relationships

The genomic evolution of symbiotic organisms, in some cases, comes along with recombination. For instance, the study of aphid endosymbiont Arsenophonus sp. revealed multiple recombination events, even when examining three genes only, indicating extensive genetic exchanges [39]. Even in obligate symbionts, Blochmannia and Buchnera, indirect evidence of recombination was found; however, they probably refer to ancestral exchanges, enabling lineage adaptation for particular hosts [20,21]. It is considered that recombination is one of the primary mechanisms of genomic evolution for Bradyrhizobium sp. and Sinorhizobium sp., as it predominantly affects symbiotic regions [40,41,42]. Recombination modulated genetic differentiation in the shipworm endosymbiont Teredinibacter turnerae [43]. Finally, Wolbachia sp., an endosymbiont of insects, is the most well-studied species in terms of recombination, and has been reported as exerting an effect on both housekeeping genes and loci encoding surface proteins performing contact with host cells [44,45,46]. There is a noticeable pattern in symbiosis-related recombination: the more stable and longer the host-symbiont relationships are, the lower the recombination intensity is. Conversely, an intensive genetic exchange appears to be a characteristic of recently adapted symbionts or those with a wide range of affected hosts.

2.3. Pathogenesis

It is not surprising that the overwhelming majority of reported inferences associated with recombination in genomes belong to pathogenic bacteria due to their importance for public health, because it provokes the origin of aggressive isolates causing infectious outbreaks involving Legionella pneumophila [47], enteropathogenic Escherichia coli O104:H4 [48], Listeria monocytogenes [49], and Neisseria meningitidis [10,11]. Notably, in some N. meningitides strains, a conspicuous HR rate was triggered by horizontal gene transfer [50]. Recombination was demonstrated orchestrating the evolution and genetic differentiation of multiple pathogenic strains relating to diverse bacteria, namely, Leptospira sp. [51], Orientia tsutsugamushi [52], Streptococcus pneumonia [53], Vibrio vulnificus [54], Yersinia pseudotuberculosis [55], Pseudomonas aeruginosa [56], Vibrio cholerae [57], Campylobacter jejuni [58], Treponema sp. [59], and Ochrobactrum intermedium [60].
Recombination mediates host adaptation during Streptococcus dysgalactiae’s evolutionary history [61]; furthermore, even a single event could lead to alterations in antigenic properties, leading to novel invasive strains [2]. Gene transfer has sufficiently contributed to the evolution of Salmonella enterica subspecies, allowing them to infect warm-blooded vertebrates [62]. It has also made the course of infection by serovars Paratyphi A and Typhi remarkably similar, even though they had been initially evolutionarily unrelated [63,64]. Clade divergence in Chlamydia trachomatis, corresponding to different infected tissues, has been accompanied by changes in the recombination level [65]. Recombination was also shown to be a major determinant of population structure during the onset of Clostridium perfringens infection [66]. Finally, intensive recombination guided the evolution of intracellular parasites Rickettsia sp. [67], but the frequency of the exchanges is lowered in more specialized lineages [68]. In some Helicobacter pylori strains, the recombination rate exceeded mutations by almost 100 times [69]. Usually, new strains tend to occur during mixed infections, increasing the probability of recombination [70] and concomitant genome rearrangements [71].
Recombination ramifications were also studied in animal and plant pathogens, in which genetic exchange forms new strains, as was reported for Bartonella sp., which infects voles [72], bats [73], and even humans [74]. Similarly, recombination governed strain divergence in the salmon pathogen Flavobacterium psychrophilum [75], the septicemic agent in pigs Streptococcus suis [76], and cattle pathogen Ehrlichia ruminantium [77,78]. Finally, recombination induces adaptation to new hosts and virulence changes in plant pathogens, leading to severe economic loss. For instance, recombination represents a genome plasticity determinant in Xanthomonas citri [79], X. gardneri [80], X. perforans [81,82], and Xylella fastidiosa [83,84], which infect citrus fruits, tomato, pepper, and coffee species, respectively. Another important example of host adaptation through recombination involves bacteria of the B. thuringiensis species. Recombination within the genes coding Cry toxins might play an important role in their diversification and adaptation to new species of insect hosts [85,86,87].

3. The Distribution of Recombination Events among the Bacterial Genome

Although virtually all parts of the bacterial genome, including conservative housekeeping genes, are more or less subjected to recombination, there are usually hotspots involved in virulence or antibiotic resistance. A bacterial genome typically contains one or several chromosomes and extrachromosomal plasmids. Recombination can affect both, as they can house mobile genetic elements (insertions, integrative elements, genomic islands) that are transferred during HGT, further causing HR due to the acquisition of homologous regions with high similarity. In the following section, we primarily discuss the role of mobile genetic elements or tandem repeats distributed among the genome, whereas the formation of new alleles and chimeric sequences in the context of certain genes is described in the next section.

3.1. Plasmids

Plasmids are extrachromosomal self-replicating genetic elements, the acquisition of which during HGT can induce antibiotic resistance or modulate virulence. Genes located on plasmids are frequently engaged in HR. This has been detected in Sinorhizobium meliloti pSymB and pSymA plasmids carrying symbiotic genes [88], the Staphylococcus aureus pWBG731 plasmid associated with methicillin resistance [89], the megaplasmid of MDR (multidrug-resistance) Pseudomonas aeruginosa strains [90], the IncF plasmid of E. coli carrying beta-lactamase [91], and the cp32/cp18 plasmids of the Lyme disease-causing agent Borrelia sp. which encode surface-located virulence factors [92,93,94]. Notably, strain B31 of Borrelia burgdorferi possesses 12 linear and 9 circular plasmids that exhibit conspicuous evidence of recombination, while the virulence potential descends both from combinations of these plasmids and the recombination within them [95].
Recombination can give birth to novel plasmid types, as shown for the virulence plasmid pAV2 of Acinetobacter baumanni [6]; in addition, chimeric plasmids with amino acid synthesis genes were found in symbiotic Buchnera aphidicola [20]. Hybrid nature is a typical trait of B. thuringiensis toxin-bearing plasmids, such as the camelysin-coding pBMB165 plasmid; pIS56-63 and pBMB0228 Cry toxins-encoding plasmids [96,97]; pAP258 and pAO254 with genes encoding for Vip and Cry toxins [98]. Occasionally, mobile genetic elements can be inserted into a plasmid, which was shown for the vancomycin-resistant Enterococcus faecium phenotype possessing Tn1546 transposons on their plasmids [99,100]. Interestingly, in Klebsiella pneumonia, plasmid-located insertion sequences caused the incorporation of beta-lactamase genes in the chromosomal genome with further excision of the insertions by homologous recombination [101]. The enterohemorrhagic E. coli O26:H-EHEC strain contains a complex plasmid pO26-CRL115, simultaneously carrying class 1 atypical integron and two transposons, Tn6026 and Tn21, and each element of the plasmid took part in extensive recombination [102]. Finally, the virulence plasmid pSDVu of Salmonella enterica serovar Dublin is considered a source for shortened virulence plasmids belonging to other S. enterica strains, while gene elimination during plasmid evolution was associated with recombination [103,104].

3.2. Insertion Sequences

Insertions and transposons belong to small mobile elements bearing the transposase gene and are capable of site-specific recombination. Apart from carrying out HGT itself, insertion sequences (IS) often initiate homologous recombination after integrating into a genome, as was demonstrated in the IS-flanked MDR region of Corynebacterium striatum [9] or the virulence factors-coding loci flanked by IS1126 and IS1272 of Porphyromonas gingivalis and Staphylococcus haemolyticus, respectively [105,106]. Insertion acquisition is linked with the emergence of new strains in Enterococcus faecium [107] and Klebsiella pneumonia [108], and in the latter, further homologous recombination promoted diversification [109]. In Clostridium botulinum, neurotoxin-encoding genes are associated with insertions, which mediate inter-strain recombination [110]. Occasionally, insertion-driven recombination may lead to large chromosomal inversions, as for E. coli O157 isolate [111], or other genome rearrangements, like gene loss and prophage elimination detected in Burkholderia mallei [112].

3.3. Long Genomic Regions

More often, yet not exclusively, recombination affects bacterial chromosomes in a gene-scale manner, although it could have an impact on large genome regions, almost inevitably producing new aggressive strains [113]. The consequences of such instances are illustrated by the examples of Vibrio parahaemolyticus outbreaks that occurred after large genome rearrangements around O- and K-antigenic regions [113], hypervirulent strains of K. pneumonia that underwent 100 kb-long exchanges [114], or strains of Legionella pneumophila and Streptococcus agalactiae, subjected to similar events throughout their evolutionary history [115,116]. In Staphylococcus aureus, such recombination-induced rearrangements cause not only hybrid strain formation [117], but also the extension of infectious potential by gaining the ability to colonize ruminants [118].

3.4. Repeats

In some microorganisms, genetic plasticity maintained by recombination is connected to short genome repeats. Tandem repeat Bams30-mediated recombination led to alterations in the exosporium composition and structure in Bacillus anthracis, and similar genomic changes contributed to anthrax-like symptoms in several B. cereus strains [119]. Bacteria with an extremely compact genome can exploit repeats as a central force for genetic evolution, like Mycoplasma pneumonia, which possesses one of the smallest genomes, yet exhibits recombination signals around repeats [120].

3.5. Genomic Islands and Integrative Elements

Genomic islands are horizontally acquired loci of varying lengths embedded into the chromosome. In a broad sense, this term describes staphylococcal chromosomal cassettes and integrative and conjugative elements (ICEs) that exploit the bacterial conjugative apparatus for self-replication [121]. Similar to insertions, these elements often flank antibiotic resistance or virulence genes and promote recombination, as was observed in Vibrio cholera ICEVchInd5 [26], Streptococcus pneumonia Tn5253-containing ICE [122], and Staphylococcus sciuri SCCmec III (staphylococcal cassette chromosome element) [123]. Special AICEs (actinomycete integrative and conjugative elements) in Streptomyces sp. ensure diversity of antimicrobial compounds through active genetic exchanges [38].
Pathogenicity islands represent one of the main targets for modern clinical microbiology because their transmittance and acquisition by bacteria induce the development of new pathogens. Moreover, the acquisition of these islands is correlated with an increased recombination rate, thus providing adaptation to hosts. Pathogenicity islands with concomitant traces of recombination were identified in the plant pathogens Acidovorax avenae [28] and Pseudomonas viridiflava [124] and numerous human infectious agents and opportunistic pathogens, including Streptococcus suis [125], Clostridioides difficile [126], Actinobacillus actinomycetemcomitans [127], E. coli [128], and Pseudomonas aeruginosa [129]. Moreover, symbiotic islands (SI) can encode symbiotic genes required for making contact with the host, and genes within them are prone to adaptive recombination, as in the case of Bradyrhizobium sp. [41].

3.6. Prophages

Prophages often serve as a means of horizontal gene transfer, and their acquisition is associated with subsequent recombination, which is commonly associated with antibiotic resistance such as in Clostridioides difficile [27] and Salmonella enterica var Typhimurium [130], or virulence as in Streptococcus pneumoniae [131]. Finally, Bartonella sp. bears demystified bacteriophages called GTAs (Gene transfer agents), the assembly of which is controlled by the host; thus, bacteria can utilize GTAs to amplify genes essential for infection [132].

4. Functional Characteristics of Genes Subjected to Recombination

4.1. Surface Proteins and Adhesion Factors

Membrane proteins are required to establish contact with host cells, being crucial for parasitic and symbiotic relationships at the onset and/or after establishment, biofilm formation, and antigenic properties. The diversity of these proteins accounts for the varieties of both virulence and symbiotic potential. Evidence of recombination is detected in Wolbachia sp. genes encoding surface proteins, namely Wsp [133] and ankyrin-rich (ANK) [45], in particular, cidA and cidB, which control cytoplasmic incompatibility [46]. Genetic diversification of Borrelia sp. with diverse infection strength is linked to recombination-mediated variations in membrane proteins OspA и OspB [92], surface antigen EppA [134], and OspE/F-like lipoproteins [135]. Similar hypervariability patterns are characteristics of OmpA and pmpE/F/H in Chlamydia trachomatis [136,137,138], incA in Chlamydia pneumonia [139], Mycoplasma pneumoniae P1 adhesin [120], Helicobacter pylori babA and babB proteins [140], Listeria monocytogenes internalins [141], Moraxella catarrhalis UspA1 and UspA2 proteins [142], pilus proteins and pspA in Streptococcus pneumoniae [143,144], Acinetobacter baumanni adhesins [145], Opa in Neisseria meningitidis [146] and membrane proteins in Legionella pneumophila [147], Staphylococcus aureus [148], and Streptococcus agalactiae [149].
Genes encoding surface proteins often have a mosaic structure, which assists in evading host immunity. This recombination-emanated mosaicism was found in fibrillar anti-phagocytic Streptococcus pyogenes proteins [150], immunoglobulin-like ligA, ligB, ligC of Leptospira sp. [151], autolysin and spA of Staphylococcus aureus [152,153], and surface lipoproteins fHbp of Neisseria meningitides [154]. In addition, in N. meningitides, an intensified recombination rate was induced shortly after the horizontal acquisition of the pgl locus, regulating glycosylation patterns; thus, rapid antigenic switches were observed [50]. In Treponema sp., the variety of antigens is delineated by the repeat-associated recombination in the tpr gene family [155].

4.2. Secretion Systems

Bacterial secretion systems, comprehensive complexes of membrane proteins, carry out the secretion of infectious effectors during the infection course. Therefore, it is no wonder that allelic diversity in the respective loci modulates virulence potential. Most of all, genes encoding type III secretion systems in Chlamydia trachomatis [65] and plant pathogens Pseudomonas syringae [156] and P. viridiflava [124] have been affected by recombination.

4.3. Infection Effectors

Effectors are typically represented by secreted proteins that control specific infectious stages, especially for intracellular pathogens. Similar to toxins, allelic diversity in respective genes may promote adaptation to particular hosts. Recombination has affected the sidJ and DotA effectors, which provide vesicle stability of Legionella pneumophila inside host phagocytes [157,158], transferring receptors of Mannheimia haemolytica [159] and Neisseria meningitidis [160], which modulate the infection course via ferric ion influx, and secretive glycosyltransferases of Streptococcus salivarius that produce a matrix for dental plaques [161].

4.4. Toxins

Allelic diversity in toxin-coding genes entails alterations in virulence or the course of infection. Thus, it is likely to be seen in genes encoding toxic compounds in multiple human pathogens, including Clostridium botulinum neurotoxins [162,163], Clostridioides difficile TcdA and TcdB (toxins A and B) [164], Mannheimia haemolytica leucotoxins [165], Streptococcus pyogenes streptolysin [166], E. coli enterotoxins [167], and Helicobacter pylori CagA oncoprotein [168]. In B. thuringiensis, insecticidal potential is maintained either by plasmid recombination, dovetailing diverse combinations of insecticidal moieties in different strains, or probable intra-gene recombination, leading to domain swapping between Cry toxins [85,86,87].

4.5. Antibiotic Resistance Genes

The development of antibiotic resistance is usually caused by the acquisition of mobile genetic elements associated with resistance genes. Obtained loci tend to recombine and mutate more frequently, hampering the prolonged application of antibiotics within one chemical class. Notably, some organisms, such as Acinetobacter baumanni, can become a reservoir of resistance for other bacteria, e.g., enterobacteria [169]. Events related to plasmids, ICEs, and insertions were reviewed above; nevertheless, recombination is found in single chromosomal genes. Examples include loci encoding permeases and transcriptional factors linked with tigecycline insensitivity in Acinetobacter baumannii [170], genes ensuring resistance to beta-lactams in Burkholderia multivorans [171], and housekeeping genes gyrA and parC associated with fluoroquinolone resistance in E.coli ST1193 clone [172].

4.6. Polysaccharide Synthesis

Outer membrane lipopolysaccharides and capsular polysaccharides act as prominent antigens for immune cells and vaccines; thus, recombination-induced variability causing serotype switching is considered to be a crucial problem in medicine, requiring the development of novel approaches in vaccine production. Recombination hotspots were located in loci controlling polysaccharide and lipopolysaccharide biosynthesis in Enterococcus faecium [173], E. coli. [174], Klebsiella pneumoniae [174], Helicobacter pylori [175], and Legionella pneumophila [147]. Importantly, these events lead to the formation of so-called clinical strains, causing nosocomial infection [173]. Recombination-maintained capsular switching, responsible for immune evasion, was observed in many pathogens, including Acinetobacter baumanni [176], Neisseria meningitidis [177], K. pneumoniae [174], Streptococcus agalactiae [149,178,179], S. pneumoniae [180], S. pyogenes [166], and the E. coli ST131 strain [181].

4.7. Metabolic Pathways

Recombination in loci encoding components of the metabolic pathways can enhance adaptation by acquiring the ability to catabolize new energy sources or by modulating the efficacy and speed of the involved enzymes. For example, recombination caused duplication and subsequent diversification of naphthalene metabolic gene clusters in Pseudomonas stutzeri [13]. Signals of homologous recombination were also revealed in the cellobiohydrolase gene in Teredinibacter turnerae [43], or gene clusters for nicotine and carbohydrates catabolism located on the horizontally acquired pAO1 megaplasmid in Arthrobacter nicotinovorans [35]. These respective genetic changes provided an adaptive advantage via expanding the metabolic potential in soil and host microenvironments. Finally, recombination affected the polyketide biosynthesis gene cluster in Streptomyces sp., thus enabling it to compete more efficiently with other bacteria by synthesizing novel antimicrobials [182]. Similar recombination-driven alterations were found in genes encoding signaling Nod-factors synthesis enzymes in Sinorhizobium sp. regulating molecular interactions with novel hosts when establishing symbiotic relationships [42].

5. Conclusions

Recombination-driven gene transfer directs prokaryotic evolution as a fundamental mechanism of genetic exchange. Having reviewed current genomic studies, we found that recombination hotspots tend to be located near horizontally acquired genes, mobile elements, and genome repeats (Figure 1a). These respective events could be classified according to three major trends, namely (i) ecological diversification, (ii) pathogenesis, and (iii) symbiosis. These exchanges mostly affect genes encoding membrane proteins, toxins, antibiotic resistance factors, polysaccharide biosynthesis enzymes, effectors, secretion systems, and metabolic pathways, with only the latter being involved in all three evolutionary trends mentioned (Figure 1b). Presumably, when microorganisms occupy a new environment, genetic exchanges between them and concomitant rearrangements occur more frequently. For instance, recombination occurs at the initial stage of adaptation to new hosts or denotes the continuous process of diversification, as in the case of Cry toxins in B. thuringiensis. Nevertheless, over time, relationships with the host become increasingly specialized, accompanied by an incremental decrease in recombination rate. Therefore, the signals of recombination are primarily detected in loci associated with environmental interaction and can be used as genomic markers displaying the ecological history of the studied strains.
In the reviewed studies, predominant methods for detecting recombination events were based on computational predictions in the genomic data. However, despite a large arsenal of tools available, current methods are not free from limitations, e.g., dependence on the models and dubious assumptions that the core genome reflects clonal relationships [5]. Nevertheless, even with possible artefactual inferences, the data from current genomic studies provided similar functional groups of genes to those reported in laboratory studies. This indicates that computational pipelines seem to correctly display the evolutionary dynamics of bacterial genomes in the context of recombination. It is noteworthy that, when analyzing the functional role of genetic exchanges, we did not select reports with predetermined criteria of grouping, however, we did reveal general trends. Undoubtedly, there is a skew toward pathogenic bacteria due to threats to global health; thus, there is a need for further studies on species that occupy other ecological niches.
Observed genomic regions with increased recombination rates might serve as a roadmap for further studies. Possible implications could include targeted analysis of genomic loci housed in mobile genetic elements, such as genomic islands or those flanked by insertions. Another strategy lies in picking particular genes, e.g., toxins of secretion systems with the reconstruction of recombination events, both ancestral and recent. Subsequent matching of these inferences with species’ phylogeny would help to reveal key adaptation steps to novel environments and identify common and/or distinct pathways within different ecological groups. All things considered, there is a strong demand for (i) performing comparative studies of recombination intensity across different bacterial species, (ii) developing new mathematical models and novel bioinformatic tools for recombination detection, and (iii) carrying out experimental validation of computationally derived observations to yield new insights and deepen our understanding of an intricate network of recombination events with their functional ramifications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/toxins15090568/s1, Table S1: Functional impact of recombination events summarizing organism, recombination-affected genome parts (genes, genome regions, mobile genetic elements, extrachromosomal parts), and functional role of the events.

Author Contributions

Conceptualization, A.E.S. and K.S.A.; formal analysis, A.E.S. and I.A.S.; project administration, A.A.N. and K.S.A.; funding acquisition, A.A.N.; writing—original draft preparation, A.E.S. and K.S.A.; writing—review and editing, A.E.S., I.A.S., A.A.N. and K.S.A.; visualization, K.S.A. and A.E.S. All authors have read and agreed to the published version of the manuscript.

Funding

The article was made with the support of the Ministry of Science and Higher Education of the Russian Federation in accordance with agreement No. 075-15-2021-1055 date 28 September 2021 on providing a grant in the form of subsidies from the Federal budget of the Russian Federation. The grant was provided for the implementation of the project: “Mobilization of the genetic resources of microorganisms on the basis of the Russian Collection of Agricultural Microorganisms (RCAM) at the All-Russia Research Institute for Agricultural Microbiology (ARRIAM) according to the network principle of organization”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We express our gratitude to the illustrator Anastasia Kureneva for the design of Figure 1.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

HGTHorizontal gene transfer
LGTLateral gene transfer
HRHomologous recombination
ICEsIntegrative and conjugative elements
MGEsMobile genetic elements
SSBSingle-stranded DNA-binding protein
SCCmecStaphylococcal cassette chromosome element
AICEsActinomycete integrative and conjugative elements
SISymbiotic island
GTAsGene transfer agents
GIsGenomic islands

References

  1. Ivars-Martínez, E.; D’Auria, G.; Rodríguez-Valera, F.; Sánchez-Porro, C.; Ventosa, A.; Joint, I.; Mühling, M. Biogeography of the Ubiquitous Marine Bacterium Alteromonas macleodii Determined by Multilocus Sequence Analysis. Mol. Ecol. 2008, 17, 4092–4106. [Google Scholar] [CrossRef] [PubMed]
  2. Chochua, S.; Rivers, J.; Mathis, S.; Li, Z.; Velusamy, S.; McGee, L.; Van Beneden, C.; Li, Y.; Metcalf, B.J.; Beall, B. Emergent Invasive Group A Streptococcus dysgalactiae subsp. equisimilis, United States, 2015–2018. Emerg. Infect. Dis. 2019, 25, 1543–1547. [Google Scholar] [CrossRef]
  3. Emamalipour, M.; Seidi, K.; Zununi Vahed, S.; Jahanban-Esfahlan, A.; Jaymand, M.; Majdi, H.; Amoozgar, Z.; Chitkushev, L.T.; Javaheri, T.; Jahanban-Esfahlan, R.; et al. Horizontal Gene Transfer: From Evolutionary Flexibility to Disease Progression. Front. Cell Dev. Biol. 2020, 8, 229. [Google Scholar] [CrossRef] [PubMed]
  4. Levin, B.R.; Cornejo, O.E. The Population and Evolutionary Dynamics of Homologous Gene Recombination in Bacteria. PLoS Genet. 2009, 5, e1000601. [Google Scholar] [CrossRef]
  5. Shikov, A.E.; Malovichko, Y.V.; Nizhnikov, A.A.; Antonets, K.S. Current Methods for Recombination Detection in Bacteria. Int. J. Mol. Sci. 2022, 23, 6257. [Google Scholar] [CrossRef] [PubMed]
  6. Fondi, M.; Bacci, G.; Brilli, M.; Papaleo, M.C.; Mengoni, A.; Vaneechoutte, M.; Dijkshoorn, L.; Fani, R. Exploring the Evolutionary Dynamics of Plasmids: The Acinetobacter Pan-Plasmidome. BMC Evol. Biol. 2010, 10, 59. [Google Scholar] [CrossRef]
  7. Blakely, G.W. Mechanisms of Horizontal Gene Transfer and DNA Recombination; Academic Press: Boston, MA, USA, 2015; Volume 1–3, ISBN 9780123971692. [Google Scholar]
  8. Wang, J.; Li, Y.; Pinto-Tomás, A.A.; Cheng, K.; Huang, Y. Habitat Adaptation Drives Speciation of a Streptomyces Species with Distinct Habitats and Disparate Geographic Origins. mBio 2022, 13, e0278121. [Google Scholar] [CrossRef]
  9. Nudel, K.; Zhao, X.; Basu, S.; Dong, X.; Hoffmann, M.; Feldgarden, M.; Allard, M.; Klompas, M.; Bry, L. Genomics of Corynebacterium striatum, an Emerging Multidrug-Resistant Pathogen of Immunocompromised Patients. Clin. Microbiol. Infect. 2018, 24, 1016.e7–1016.e13. [Google Scholar] [CrossRef]
  10. Hao, L.; Holden, M.T.G.; Wang, X.; Andrew, L.; Wellnitz, S.; Hu, F.; Whaley, M.; Sammons, S.; Knipe, K.; Frace, M.; et al. Distinct Evolutionary Patterns of Neisseria meningitidis Serogroup B Disease Outbreaks at Two Universities in the USA. Microb. Genom. 2018, 4, e000155. [Google Scholar] [CrossRef]
  11. Guo, Q.; Mustapha, M.M.; Chen, M.; Qu, D.; Zhang, X.; Chen, M.; Doi, Y.; Wang, M.; Harrison, L.H. Evolution of Sequence Type 4821 Clonal Complex Meningococcal Strains in China from Prequinolone to Quinolone Era, 1972–2013. Emerg. Infect. Dis. 2018, 24, 683–690. [Google Scholar] [CrossRef]
  12. Tong, W.; Li, X.; Wang, E.; Cao, Y.; Chen, W.; Tao, S.; Wei, G. Genomic Insight into the Origins and Evolution of Symbiosis Genes in Phaseolus vulgaris Microsymbionts. BMC Genom. 2020, 21, 186. [Google Scholar] [CrossRef]
  13. Bosch, R.; GarcıÍa-Valdés, E.; Moore, E.R.B. Complete Nucleotide Sequence and Evolutionary Significance of a Chromosomally Encoded Naphthalene-Degradation Lower Pathway from Pseudomonas stutzeri AN10. Gene 2000, 245, 65–74. [Google Scholar] [CrossRef] [PubMed]
  14. Didelot, X.; Maiden, M.C.J. Impact of Recombination on Bacterial Evolution. Trends Microbiol. 2010, 18, 315–322. [Google Scholar] [CrossRef] [PubMed]
  15. Ochman, H.; Lawrence, J.G.; Groisman, E.A. Lateral Gene Transfer and the Nature of Bacterial Innovation. Nature 2000, 405, 299–304. [Google Scholar] [CrossRef] [PubMed]
  16. Thomas, C.M.; Nielsen, K.M. Mechanisms of, and Barriers to, Horizontal Gene Transfer between Bacteria. Nat. Rev. Microbiol. 2005, 3, 711–721. [Google Scholar] [CrossRef]
  17. Chun, J.; Grim, C.J.; Hasan, N.A.; Lee, J.H.; Choi, S.Y.; Haley, B.J.; Taviani, E.; Jeon, Y.-S.; Kim, D.-W.; Lee, J.-H.; et al. Comparative Genomics Reveals Mechanism for Short-Term and Long-Term Clonal Transitions in Pandemic Vibrio cholerae. Proc. Natl. Acad. Sci. USA 2009, 106, 15442–15447. [Google Scholar] [CrossRef] [PubMed]
  18. Julin, D.A. Recombination: Mechanisms, Pathways, and Applications; Wells, R., Bond, J., Klinman, J., Masters, B., Bell, E., Eds.; Molecular Life Sciences; Springer: New York, NY, USA, 2017; pp. 1–28. [Google Scholar] [CrossRef]
  19. Rocha, E.P.C.; Cornet, E.; Michel, B. Comparative and Evolutionary Analysis of the Bacterial Homologous Recombination Systems. PLoS Genet. 2005, 1, e15. [Google Scholar] [CrossRef] [PubMed]
  20. Gil, R.; Sabater-Muñoz, B.; Perez-Brocal, V.; Silva, F.J.; Latorre, A. Plasmids in the Aphid Endosymbiont Buchnera aphidicola with the Smallest Genomes. A Puzzling Evolutionary Story. Gene 2006, 370, 17–25. [Google Scholar] [CrossRef]
  21. Williams, L.E.; Wernegreen, J.J. Genome Evolution in an Ancient Bacteria-Ant Symbiosis: Parallel Gene Loss among Blochmannia Spanning the Origin of the Ant Tribe Camponotini. PeerJ 2015, 3, e881. [Google Scholar] [CrossRef]
  22. Zhou, K.; Aertsen, A.; Michiels, C.W. The Role of Variable DNA Tandem Repeats in Bacterial Adaptation. FEMS Microbiol. Rev. 2014, 38, 119–141. [Google Scholar] [CrossRef]
  23. Iranzo, J.; Wolf, Y.I.; Koonin, E.V.; Sela, I. Gene Gain and Loss Push Prokaryotes beyond the Homologous Recombination Barrier and Accelerate Genome Sequence Divergence. Nat. Commun. 2019, 10, 5376. [Google Scholar] [CrossRef] [PubMed]
  24. Ely, B. Recombination and Gene Loss Occur Simultaneously during Bacterial Horizontal Gene Transfer. PLoS ONE 2020, 15, e0227987. [Google Scholar] [CrossRef] [PubMed]
  25. Gratia, J.-P. Genetic Recombinational Events in Prokaryotes and Their Viruses: Insight into the Study of Evolution and Biodiversity. Antonie Leeuwenhoek 2017, 110, 1493–1514. [Google Scholar] [CrossRef]
  26. Spagnoletti, M.; Ceccarelli, D.; Rieux, A.; Fondi, M.; Taviani, E.; Fani, R.; Colombo, M.M.; Colwell, R.R.; Balloux, F. Acquisition and Evolution of SXT-R391 Integrative Conjugative Elements in the Seventh-Pandemic Vibrio cholerae Lineage. mBio 2014, 5, e01356-14. [Google Scholar] [CrossRef] [PubMed]
  27. Garneau, J.R.; Sekulovic, O.; Dupuy, B.; Soutourina, O.; Monot, M.; Fortier, L.-C. High Prevalence and Genetic Diversity of Large PhiCD211 (PhiCDIF1296T)-Like Prophages in Clostridioides difficile. Appl. Environ. Microbiol. 2018, 84, e02164-17. [Google Scholar] [CrossRef] [PubMed]
  28. Zeng, Q.; Wang, J.; Bertels, F.; Giordano, P.R.; Chilvers, M.I.; Huntley, R.B.; Vargas, J.M.; Sundin, G.W.; Jacobs, J.L.; Yang, C.-H. Recombination of Virulence Genes in Divergent Acidovorax avenae Strains That Infect a Common Host. Mol. Plant-Microbe Interact. 2017, 30, 813–828. [Google Scholar] [CrossRef]
  29. Vos, M.; Didelot, X. A Comparison of Homologous Recombination Rates in Bacteria and Archaea. ISME J. 2009, 3, 199–208. [Google Scholar] [CrossRef]
  30. González-Torres, P.; Rodríguez-Mateos, F.; Antón, J.; Gabaldón, T. Impact of Homologous Recombination on the Evolution of Prokaryotic Core Genomes. mBio 2019, 10, e02494-18. [Google Scholar] [CrossRef]
  31. Helgason, E.; Caugant, D.A.; Lecadet, M.M.; Chen, Y.; Mahillon, J.; Lövgren, A.; Hegna, I.; Kvaløy, K.; Kolstø, A.B. Genetic Diversity of Bacillus cereus/B. thuringiensis Isolates from Natural Sources. Curr. Microbiol. 1998, 37, 80–87. [Google Scholar] [CrossRef]
  32. Liu, Q.; Xin, Y.H.; Zhou, Y.G.; Chen, W.X. Multilocus Sequence Analysis of Homologous Recombination and Diversity in Arthrobacter sensu lato Named Species and Glacier-Inhabiting Strains. Syst. Appl. Microbiol. 2018, 41, 23–29. [Google Scholar] [CrossRef]
  33. Hoetzinger, M.; Hahn, M.W. Genomic Divergence and Cohesion in a Species of Pelagic Freshwater Bacteria. BMC Genom. 2017, 18, 794. [Google Scholar] [CrossRef] [PubMed]
  34. Nesbø, C.L.; Swithers, K.S.; Dahle, H.; Haverkamp, T.H.A.; Birkeland, N.K.; Sokolova, T.; Kublanov, I.; Zhaxybayeva, O. Evidence for Extensive Gene Flow and Thermotoga Subpopulations in Subsurface and Marine Environments. ISME J. 2015, 9, 1532–1542. [Google Scholar] [CrossRef] [PubMed]
  35. Mihasan, M.; Brandsch, R. PAO1 of Arthrobacter nicotinovorans and the Spread of Catabolic Traits by Horizontal Gene Transfer in Gram-Positive Soil Bacteria. J. Mol. Evol. 2013, 77, 22–30. [Google Scholar] [CrossRef] [PubMed]
  36. Sogge, H.; Rohrlack, T.; Rounge, T.B.; Sønstebø, J.H.; Tooming-Klunderud, A.; Kristensen, T.; Jakobsena, K.S. Gene Flow, Recombination, and Selection in Cyanobacteria: Population Structure of Geographically Related Planktothrix Freshwater Strains. Appl. Environ. Microbiol. 2013, 79, 508–515. [Google Scholar] [CrossRef]
  37. Rounge, T.B.; Rohrlack, T.; Kristensen, T.; Jakobsen, K.S. Recombination and Selectional Forces in Cyanopeptolin NRPS Operons from Highly Similar, but Geographically Remote Planktothrix Strains. BMC Microbiol. 2008, 8, 141. [Google Scholar] [CrossRef]
  38. Tidjani, A.R.; Lorenzi, J.N.; Toussaint, M.; Van Dijk, E.; Naquin, D.; Lespinet, O.; Bontemps, C.; Leblond, P. Massive Gene Flux Drives Genome Diversity between Sympatric Streptomyces Conspecifics. mBio 2019, 10, e01533-19. [Google Scholar] [CrossRef]
  39. Mouton, L.; Thierry, M.; Henri, H.; Baudin, R.; Gnankine, O.; Reynaud, B.; Zchori-Fein, E.; Becker, N.; Fleury, F.; Delatte, H. Evidence of Diversity and Recombination in Arsenophonus Symbionts of the Bemisia tabaci Species Complex. BMC Microbiol. 2012, 12, S10. [Google Scholar] [CrossRef]
  40. Naamala, J.; Jaiswal, S.K.; Dakora, F.D. Microsymbiont Diversity and Phylogeny of Native Bradyrhizobia Associated with Soybean (Glycine max L. Merr.) Nodulation in South African Soils. Syst. Appl. Microbiol. 2016, 39, 336–344. [Google Scholar] [CrossRef]
  41. Parker, M.A. Legumes Select Symbiosis Island Sequence Variants in Bradyrhizobium. Mol. Ecol. 2012, 21, 1769–1778. [Google Scholar] [CrossRef]
  42. Guo, H.J.; Wang, E.T.; Zhang, X.X.; Li, Q.Q.; Zhang, Y.M.; Tian, C.F.; Chen, W.X. Replicon-Dependent Differentiation of Symbiosis-Related Genes in Sinorhizobium Strains Nodulating Glycine Max. Appl. Environ. Microbiol. 2014, 80, 1245–1255. [Google Scholar] [CrossRef]
  43. Altamia, M.A.; Wood, N.; Fung, J.M.; Dedrick, S.; Linton, E.W.; Concepcion, G.P.; Haygood, M.G.; Distel, D.L. Genetic Differentiation among Isolates of Teredinibacter turnerae, a Widely Occurring Intracellular Endosymbiont of Shipworms. Mol. Ecol. 2014, 23, 1418–1432. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, X.; Xiong, X.; Cao, W.; Zhang, C.; Werren, J.H.; Wang, X. Phylogenomic Analysis of Wolbachia Strains Reveals Patterns of Genome Evolution and Recombination. Genome Biol. Evol. 2020, 12, 2508–2520. [Google Scholar] [CrossRef] [PubMed]
  45. Siozios, S.; Ioannidis, P.; Klasson, L.; Andersson, S.G.E.; Braig, H.R.; Bourtzis, K. The Diversity and Evolution of Wolbachia Ankyrin Repeat Domain Genes. PLoS ONE 2013, 8, e55390. [Google Scholar] [CrossRef] [PubMed]
  46. Bonneau, M.; Atyame, C.; Beji, M.; Justy, F.; Cohen-Gonsaud, M.; Sicard, M.; Weill, M. Culex pipiens Crossing Type Diversity Is Governed by an Amplified and Polymorphic Operon of Wolbachia. Nat. Commun. 2018, 9, 319. [Google Scholar] [CrossRef]
  47. Sánchez-Busó, L.; Comas, I.; Jorques, G.; González-Candelas, F. Recombination Drives Genome Evolution in Outbreak-Related Legionella pneumophila Isolates. Nat. Genet. 2014, 46, 1205–1211. [Google Scholar] [CrossRef]
  48. Hao, W.; Allen, V.G.; Jamieson, F.B.; Low, D.E.; Alexander, D.C. Phylogenetic Incongruence in E. coli O104: Understanding the Evolutionary Relationships of Emerging Pathogens in the Face of Homologous Recombination. PLoS ONE 2012, 7, e33971. [Google Scholar] [CrossRef]
  49. Orsi, R.H.; Borowsky, M.L.; Lauer, P.; Young, S.K.; Nusbaum, C.; Galagan, J.E.; Birren, B.W.; Ivy, R.A.; Sun, Q.; Graves, L.M.; et al. Short-term Genome Evolution of Listeria monocytogenes in a Non-Controlled Environment. BMC Genom. 2008, 9, 539. [Google Scholar] [CrossRef]
  50. Lamelas, A.; Harris, S.R.; Röltgen, K.; Dangy, J.P.; Hauser, J.; Kingsley, R.A.; Connor, T.R.; Sie, A.; Hodgson, A.; Dougan, G.; et al. Emergence of a New Epidemic Neisseria meningitidis Serogroup a Clone in the African Meningitis Belt: High-Resolution Picture of Genomic Changes That Mediate Immune Evasion. mBio 2014, 5, e01974-14. [Google Scholar] [CrossRef]
  51. Caimi, K.; Repetto, S.A.; Varni, V.; Ruybal, P. Leptospira Species Molecular Epidemiology in the Genomic Era. Infect. Genet. Evol. 2017, 54, 478–485. [Google Scholar] [CrossRef]
  52. Kim, G.; Ha, N.Y.; Min, C.K.; Kim, H.I.; Yen, N.T.H.; Lee, K.H.; Oh, I.; Kang, J.S.; Choi, M.S.; Kim, I.S.; et al. Diversification of Orientia tsutsugamushi Genotypes by Intragenic Recombination and Their Potential Expansion in Endemic Areas. PLoS Neglected Trop. Dis. 2017, 11, e5408. [Google Scholar] [CrossRef]
  53. Donati, C.; Hiller, N.L.; Tettelin, H.; Muzzi, A.; Croucher, N.J.; Angiuoli, S.V.; Oggioni, M.; Dunning Hotopp, J.C.; Hu, F.Z.; Riley, D.R.; et al. Structure and Dynamics of the Pan-Genome of Streptococcus pneumoniae and Closely Related Species. Genome Biol. 2010, 11, R107. [Google Scholar] [CrossRef] [PubMed]
  54. Bisharat, N.; Cohen, D.I.; Maiden, M.C.; Crook, D.W.; Peto, T.; Harding, R.M. The Evolution of Genetic Structure in the Marine Pathogen, Vibrio vulnificus. Infect. Genet. Evol. 2007, 7, 685–693. [Google Scholar] [CrossRef] [PubMed]
  55. Ch’Ng, S.L.; Octavia, S.; Xia, Q.; Duong, A.; Tanaka, M.M.; Fukushima, H.; Lan, R. Population Structure and Evolution of Pathogenicity of Yersinia pseudotuberculosis. Appl. Environ. Microbiol. 2011, 77, 768–775. [Google Scholar] [CrossRef]
  56. Waine, D.J.; Honeybourne, D.; Smith, E.G.; Whitehouse, J.L.; Dowson, C.G. Cross-Sectional and Longitudinal Multilocus Sequence Typing of Pseudomonas aeruginosa in Cystic Fibrosis Sputum Samples. J. Clin. Microbiol. 2009, 47, 3444–3448. [Google Scholar] [CrossRef]
  57. Feng, L.; Reeves, P.R.; Lan, R.; Ren, Y.; Gao, C.; Zhou, Z.; Ren, Y.; Cheng, J.; Wang, W.; Wang, J.; et al. A Recalibrated Molecular Clock and Independent Origins for the Cholera Pandemic Clones. PLoS ONE 2008, 3, e4053. [Google Scholar] [CrossRef]
  58. Biggs, P.J.; Fearnhead, P.; Hotter, G.; Mohan, V.; Collins-Emerson, J.; Kwan, E.; Besser, T.E.; Cookson, A.; Carter, P.E.; French, N.P. Whole-Genome Comparison of Two Campylobacter jejuni Isolates of the Same Sequence Type Reveals Multiple Loci of Different Ancestral Lineage. PLoS ONE 2011, 6, e27121. [Google Scholar] [CrossRef]
  59. Štaudová, B.; Strouhal, M.; Zobaníková, M.; Čejková, D.; Fulton, L.L.; Chen, L.; Giacani, L.; Centurion-Lara, A.; Bruisten, S.M.; Sodergren, E.; et al. Whole Genome Sequence of the Treponema pallidum subsp. endemicum Strain Bosnia A: The Genome Is Related to Yaws Treponemes but Contains Few Loci Similar to Syphilis Treponemes. PLoS Neglected Trop. Dis. 2014, 8, e3261. [Google Scholar] [CrossRef] [PubMed]
  60. Aujoulat, F.; Romano-Bertrand, S.; Masnou, A.; Marchandin, H.; Jumas-Bilak, E. Niches, Population Structure and Genome Reduction in Ochrobactrum intermedium: Clues to Technology-Driven Emergence of Pathogens. PLoS ONE 2014, 9, e83376. [Google Scholar] [CrossRef]
  61. McMillan, D.J.; Kaul, S.Y.; Bramhachari, P.V.; Smeesters, P.R.; Vu, T.; Karmarkar, M.G.; Shaila, M.S.; Sriprakash, K.S. Recombination Drives Genetic Diversification of Streptococcus dysgalactiae subspecies equisimilis in a Region of Streptococcal Endemicity. PLoS ONE 2011, 6, e21346. [Google Scholar] [CrossRef]
  62. Desai, P.T.; Porwollik, S.; Long, F.; Cheng, P.; Wollam, A.; Clifton, S.W.; Weinstock, G.M.; McClelland, M. EvolutionaryGenomics of Salmonella enterica Subspecies. mBio 2013, 4, e00579-12. [Google Scholar] [CrossRef]
  63. Didelot, X.; Achtman, M.; Parkhill, J.; Thomson, N.R.; Falush, D. A Bimodal Pattern of Relatedness between the Salmonella Paratyphi A and Typhi Genomes: Convergence or Divergence by Homologous Recombination? Genome Res. 2007, 17, 61–68. [Google Scholar] [CrossRef] [PubMed]
  64. Holt, K.E.; Thomson, N.R.; Wain, J.; Langridge, G.C.; Hasan, R.; Bhutta, Z.A.; Quail, M.A.; Norbertczak, H.; Walker, D.; Simmonds, M.; et al. Pseudogene Accumulation in the Evolutionary Histories of Salmonella enterica Serovars Paratyphi A and Typhi. BMC Genom. 2009, 10, 36. [Google Scholar] [CrossRef] [PubMed]
  65. Joseph, S.J.; Didelot, X.; Rothschild, J.; De Vries, H.J.C.; Morré, S.A.; Read, T.D.; Dean, D. Population Genomics of Chlamydia trachomatis: Insights on Drift, Selection, Recombination, and Population Structure. Mol. Biol. Evol. 2012, 29, 3933–3946. [Google Scholar] [CrossRef] [PubMed]
  66. Sawires, Y.S.; Songer, J.G. Clostridium perfringens: Insight into Virulence Evolution and Population Structure. Anaerobe 2006, 12, 23–43. [Google Scholar] [CrossRef]
  67. Wu, J.; Yu, T.; Bao, Q.; Zhao, F. Evidence of Extensive Homologous Recombination in the Core Genome of Rickettsia. Comp. Funct. Genom. 2009, 2009, 510270. [Google Scholar] [CrossRef]
  68. Hernández-López, A.; Chabrol, O.; Royer-Carenzi, M.; Merhej, V.; Pontarotti, P.; Raoult, D. To Tree or Not to Tree? Genome-Wide Quantification of Recombination and Reticulate Evolution during the Diversification of Strict Intracellular Bacteria. Genome Biol. Evol. 2013, 5, 2305–2317. [Google Scholar] [CrossRef]
  69. Didelot, X.; Nell, S.; Yang, I.; Woltemate, S.; Van Der Merwe, S.; Suerbaum, S. Genomic Evolution and Transmission of Helicobacter pylori in Two South African Families. Proc. Natl. Acad. Sci. USA 2013, 110, 13880–13885. [Google Scholar] [CrossRef]
  70. Krebes, J.; Didelot, X.; Kennemann, L.; Suerbaum, S. Bidirectional Genomic Exchange between Helicobacter pylori Strains from a Family in Coventry, United Kingdom. Int. J. Med. Microbiol. 2014, 304, 1135–1146. [Google Scholar] [CrossRef]
  71. Lara-Ramírez, E.E.; Segura-Cabrera, A.; Guo, X.; Yu, G.; García-Pérez, C.A.; Rodríguez-Pérez, M.A. New Implications on Genomic Adaptation Derived from the Helicobacter pylori Genome Comparison. PLoS ONE 2011, 6, e17300. [Google Scholar] [CrossRef]
  72. Paziewska, A.; Harris, P.D.; Zwolińska, L.; Bajer, A.; Siński, E. Recombination Within and Between Species of the Alpha Proteobacterium Bartonella Infecting Rodents. Microb. Ecol. 2011, 61, 134–145. [Google Scholar] [CrossRef]
  73. Bai, Y.; Hayman, D.T.S.; McKee, C.D.; Kosoy, M.Y. Classification of Bartonella Strains Associated with Straw-Colored Fruit Bats (Eidolon helvum) across Africa Using a Multi-Locus Sequence Typing Platform. PLoS Neglected Trop. Dis. 2015, 9, e3478. [Google Scholar] [CrossRef] [PubMed]
  74. Berglund, E.C.; Ellegaard, K.; Granberg, F.; Xie, Z.; Maruyama, S.; Kosoy, M.Y.; Birtles, R.J.; Andersson, S.G.E. Rapid Diversification by Recombination in Bartonella grahamii from Wild Rodents in Asia Contrasts with Low Levels of Genomic Divergence in Northern Europe and America. Mol. Ecol. 2010, 19, 2241–2255. [Google Scholar] [CrossRef] [PubMed]
  75. Knupp, C.; Wiens, G.D.; Faisal, M.; Call, D.R.; Cain, K.D.; Nicolas, P.; Van Vliet, D.; Yamashita, C.; Ferguson, J.A.; Meuninck, D.; et al. Large-Scale Analysis of Flavobacterium psychrophilum Multilocus Sequence Typing Genotypes Recovered from North American Salmonids Indicates That Both Newly Identified and Recurrent Clonal Complexes Are Associated with Disease. Appl. Environ. Microbiol. 2019, 85, e02305-18. [Google Scholar] [CrossRef] [PubMed]
  76. Athey, T.B.T.; Auger, J.-P.; Teatero, S.; Dumesnil, A.; Takamatsu, D.; Wasserscheid, J.; Dewar, K.; Gottschalk, M.; Fittipaldi, N. Complex Population Structure and Virulence Differences among Serotype 2 Streptococcus suis Strains Belonging to Sequence Type 28. PLoS ONE 2015, 10, e137760. [Google Scholar] [CrossRef]
  77. Cangi, N.; Gordon, J.L.; Bournez, L.; Pinarello, V.; Aprelon, R.; Huber, K.; Lefrançois, T.; Neves, L.; Meyer, D.F.; Vachiéry, N. Recombination Is a Major Driving Force of Genetic Diversity in the Anaplasmataceae Ehrlichia ruminantium. Front. Cell. Infect. Microbiol. 2016, 6, 111. [Google Scholar] [CrossRef]
  78. Adakal, H.; Meyer, D.F.; Carasco-Lacombe, C.; Pinarello, V.; Allègre, F.; Huber, K.; Stachurski, F.; Morand, S.; Martinez, D.; Lefrançois, T.; et al. MLST Scheme of Ehrlichia ruminantium: Genomic Stasis and Recombination in Strains from Burkina-Faso. Infect. Genet. Evol. 2009, 9, 1320–1328. [Google Scholar] [CrossRef]
  79. Gordon, J.L.; Lefeuvre, P.; Escalon, A.; Barbe, V.; Cruveiller, S.; Gagnevin, L.; Pruvost, O. Comparative Genomics of 43 Strains of Xanthomonas citri pv. Citri Reveals the Evolutionary Events Giving Rise to Pathotypes with Different Host Ranges. BMC Genom. 2015, 16, 1098. [Google Scholar] [CrossRef]
  80. Timilsina, S.; Jibrin, M.O.; Potnis, N.; Minsavage, G.V.; Kebede, M.; Schwartz, A.; Bart, R.; Staskawicz, B.; Boyer, C.; Vallad, G.E.; et al. Multilocus Sequence Analysis of Xanthomonads Causing Bacterial Spot of Tomato and Pepper Plants Reveals Strains Generated by Recombination among Species and Recent Global Spread of Xanthomonas gardneri. Appl. Environ. Microbiol. 2015, 81, 1520–1529. [Google Scholar] [CrossRef]
  81. Newberry, E.A.; Bhandari, R.; Minsavage, G.V.; Timilsina, S.; Jibrin, M.O.; Kemble, J.; Sikora, E.J.; Jones, J.B.; Potnis, N. Independent Evolution with the Gene Flux Originating from Multiple Xanthomonas Species Explains Genomic Heterogeneity in Xanthomonas perforans. Appl. Environ. Microbiol. 2019, 85, e00885-19. [Google Scholar] [CrossRef]
  82. Timilsina, S.; Pereira-Martin, J.A.; Minsavage, G.V.; Iruegas-Bocardo, F.; Abrahamian, P.; Potnis, N.; Kolaczkowski, B.; Vallad, G.E.; Goss, E.M.; Jones, J.B. Multiple Recombination Events Drive the Current Genetic Structure of Xanthomonas perforans in Florida. Front. Microbiol. 2019, 10, 448. [Google Scholar] [CrossRef]
  83. Potnis, N.; Kandel, P.P.; Merfa, M.V.; Retchless, A.C.; Parker, J.K.; Stenger, D.C.; Almeida, R.P.P.; Bergsma-Vlami, M.; Westenberg, M.; Cobine, P.A.; et al. Patterns of Inter- and Intrasubspecific Homologous Recombination Inform Eco-Evolutionary Dynamics of Xylella fastidiosa. ISME J. 2019, 13, 2319–2333. [Google Scholar] [CrossRef]
  84. Jacques, M.A.; Denancé, N.; Legendre, B.; Morel, E.; Briand, M.; Mississipi, S.; Durand, K.; Olivier, V.; Portier, P.; Poliakoff, F.; et al. New Coffee Plant-Infecting Xylella fastidiosa Variants Derived via Homologous Recombination. Appl. Environ. Microbiol. 2016, 82, 1556–1568. [Google Scholar] [CrossRef]
  85. Höfte, H.; De Greve, H.; Seurinck, J.; Jansens, S.; Mahillon, J.; Ampe, C.; Vandekerckhove, J.; Vanderbruggen, H.; Van Montagu, M.; Zabeau, M.; et al. Structural and Functional Analysis of a Cloned Delta Endotoxin of Bacillus thuringiensis Berliner 1715. Eur. J. Biochem. 1986, 161, 273–280. [Google Scholar] [CrossRef] [PubMed]
  86. Shu, C.; Zhang, F.; Chen, G.; Joseph, L.; Barqawi, A.; Evans, J.; Song, F.; Li, G.; Zhang, J.; Crickmore, N. A Natural Hybrid of a Bacillus thuringiensis Cry2A Toxin Implicates Domain I in Specificity Determination. J. Invertebr. Pathol. 2017, 150, 35–40. [Google Scholar] [CrossRef] [PubMed]
  87. De Maagd, R.A.; Bravo, A.; Crickmore, N. How Bacillus thuringiensis Has Evolved Specific Toxins to Colonize the Insect World. Trends Genet. 2001, 17, 193–199. [Google Scholar] [CrossRef] [PubMed]
  88. Toro, N.; Martínez-Abarca, F.; Fernández-López, M. The Early Events Underlying Genome Evolution in a Localized Sinorhizobium meliloti Population. BMC Genom. 2016, 17, 556. [Google Scholar] [CrossRef]
  89. Eto, K.Y.; Firth, N.; Davis, A.M.; Kwong, S.M.; Krysiak, M.; Lee, Y.T.; O’Brien, F.G.; Grubb, W.B.; Coombs, G.W.; Bond, C.S.; et al. Evolution of a 72-Kilobase Cointegrant, Conjugative Multiresistance Plasmid in Community-Associated Methicillin-Resistant Staphylococcus aureus Isolates from the Early 1990s. Antimicrob. Agents Chemother. 2019, 63, e01560-19. [Google Scholar] [CrossRef]
  90. Cazares, A.; Moore, M.P.; Hall, J.P.J.; Wright, L.L.; Grimes, M.; Emond-Rhéault, J.-G.; Pongchaikul, P.; Santanirand, P.; Levesque, R.C.; Fothergill, J.L.; et al. A Megaplasmid Family Driving Dissemination of Multidrug Resistance in Pseudomonas. Nat. Commun. 2020, 11, 1370. [Google Scholar] [CrossRef]
  91. Schink, A.K.; Kadlec, K.; Kaspar, H.; Mankertz, J.; Schwarz, S. Analysis of Extended-Spectrum-β-Lactamase-Producing Escherichia coli Isolates Collected in the GERM-Vet Monitoring Programme. J. Antimicrob. Chemother. 2013, 68, 1741–1749. [Google Scholar] [CrossRef]
  92. Yabuki, M.; Nakao, M.; Fukunaga, M. Genetic Diversity and the Absence of Regional Differences of Borrelia garinii as Demonstrated by ospA and ospB Gene Sequence Analysis. Microbiol. Immunol. 1999, 43, 1097–1102. [Google Scholar] [CrossRef]
  93. Taylor, J.C.; Martin, H.C.; Lise, S.; Broxholme, J.; Cazier, J.-B.; Rimmer, A.; Kanapin, A.; Lunter, G.; Fiddy, S.; Allan, C.; et al. Factors Influencing Success of Clinical Genome Sequencing across a Broad Spectrum of Disorders. Nat. Genet. 2015, 47, 717–726. [Google Scholar] [CrossRef] [PubMed]
  94. Stevenson, B.; Casjens, S.; Rosa, P. Evidence of Past Recombination Events among the Genes Encoding the Erp Antigens of Borrelia burgdorferi. Microbiology 1998, 144, 1869–1879. [Google Scholar] [CrossRef] [PubMed]
  95. Casjens, S.R.; Di, L.; Akther, S.; Mongodin, E.F.; Luft, B.J.; Schutzer, S.E.; Fraser, C.M.; Qiu, W.G. Primordial Origin and Diversification of Plasmids in Lyme Disease Agent Bacteria. BMC Genom. 2018, 19, 218. [Google Scholar] [CrossRef]
  96. Murawska, E.; Fiedoruk, K.; Swiecicka, I. Modular Genetic Architecture of the Toxigenic Plasmid pIS56-63 Harboring cry1Ab21 in Bacillus thuringiensis subsp. thuringiensis Strain IS5056. Pol. J. Microbiol. 2014, 63, 147–156. [Google Scholar] [CrossRef] [PubMed]
  97. Wang, P.; Zhang, C.; Zhu, Y.; Deng, Y.; Guo, S.; Peng, D.; Ruan, L.; Sun, M. The Resolution and Regeneration of a Cointegrate Plasmid Reveals a Model for Plasmid Evolution Mediated by Conjugation and OriT Site-Specific Recombination. Environ. Microbiol. 2013, 15, 3305–3318. [Google Scholar] [CrossRef] [PubMed]
  98. Wang, Z.; Wang, K.; Bravo, A.; Soberón, M.; Cai, J.; Shu, C.; Zhang, J. Coexistence of Cry9 with the Vip3A Gene in an Identical Plasmid of Bacillus thuringiensis Indicates Their Synergistic Insecticidal Toxicity. J. Agric. Food Chem. 2020, 68, 14081–14090. [Google Scholar] [CrossRef]
  99. Wardal, E.; Kuch, A.; Gawryszewska, I.; Żabicka, D.; Hryniewicz, W.; Sadowy, E. Diversity of Plasmids and Tn1546-Type Transposons among VanA Enterococcus faecium in Poland. Eur. J. Clin. Microbiol. Infect. Dis. 2017, 36, 313–328. [Google Scholar] [CrossRef]
  100. van Hal, S.J.; Beukers, A.G.; Timms, V.J.; Ellem, J.A.; Taylor, P.; Maley, M.W.; Newton, P.J.; Ferguson, J.K.; Lee, A.; Chen, S.C.-A.; et al. Relentless Spread and Adaptation of Non-Typeable VanA Vancomycin-resistant Enterococcus faecium: A Genome-Wide Investigation. J. Antimicrob. Chemother. 2018, 73, 1487–1491. [Google Scholar] [CrossRef]
  101. Hudson, C.M.; Bent, Z.W.; Meagher, R.J.; Williams, K.P. Resistance Determinants and Mobile Genetic Elements of an NDM-1-Encoding Klebsiella pneumoniae Strain. PLoS ONE 2014, 9, e99209. [Google Scholar] [CrossRef]
  102. Venturini, C.; Hassan, K.A.; Chowdhury, P.R.; Paulsen, I.T.; Walker, M.J.; Djordjevic, S.P. Sequences of Two Related Multiple Antibiotic Resistance Virulence Plasmids Sharing a Unique IS26-Related Molecular Signature Isolated from Different Escherichia coli Pathotypes from Different Hosts. PLoS ONE 2013, 8, e78862. [Google Scholar] [CrossRef]
  103. Hong, S.F.; Chiu, C.H.; Chu, C.; Feng, Y.; Ou, J.T. Complete Nucleotide Sequence of a Virulence Plasmid of Salmonella enterica Serovar Dublin and Its Phylogenetic Relationship to the Virulence Plasmids of Serovars Choleraesuis, Enteritidis and Typhimurium. FEMS Microbiol. Lett. 2008, 282, 39–43. [Google Scholar] [CrossRef]
  104. Chu, C.; Feng, Y.; Chien, A.C.; Hu, S.; Chu, C.H.; Chiu, C.H. Evolution of Genes on the Salmonella Virulence Plasmid Phylogeny Revealed from Sequencing of the Virulence Plasmids of S. enterica Serotype Dublin and Comparative Analysis. Genomics 2008, 92, 339–343. [Google Scholar] [CrossRef]
  105. Dong, H.; Chen, T.; Dewhirst, F.E.; Fleischmann, R.D.; Fraser, C.M.; Duncan, M.J. Genomic Loci of the Porphyromonas gingivalis Insertion Element IS 1126. Infect. Immun. 1999, 67, 3416–3423. [Google Scholar] [CrossRef]
  106. Bouchami, O.; De Lencastre, H.; Miragaia, M. Impact of Insertion Sequences and Recombination on the Population Structure of Staphylococcus haemolyticus. PLoS ONE 2016, 11, e0156653. [Google Scholar] [CrossRef] [PubMed]
  107. Bayjanov, J.R.; Baan, J.; Rogers, M.R.C.; Troelstra, A.; Willems, R.J.L.; van Schaik, W. Enterococcus faecium Genome Dynamics during Long-Term Asymptomatic Patient Gut Colonization. Microb. Genom. 2019, 5, e000277. [Google Scholar] [CrossRef] [PubMed]
  108. Wyres, K.L.; Gorrie, C.; Edwards, D.J.; Wertheim, H.F.L.; Hsu, L.Y.; Van Kinh, N.; Zadoks, R.; Baker, S.; Holt, K.E. Extensive Capsule Locus Variation and Large-Scale Genomic Recombination within the Klebsiella pneumoniae Clonal Group 258. Genome Biol. Evol. 2015, 7, 1267–1279. [Google Scholar] [CrossRef] [PubMed]
  109. Chen, L.; Mathema, B.; Pitout, J.D.D.; DeLeo, F.R.; Kreiswirth, B.N. Epidemic Klebsiella pneumoniae ST258 Is a Hybrid Strain. mBio 2014, 5, e01355-14. [Google Scholar] [CrossRef]
  110. Hill, K.K.; Xie, G.; Foley, B.T.; Smith, T.J.; Munk, A.C.; Bruce, D.; Smith, L.A.; Brettin, T.S.; Detter, J.C. Recombination and Insertion Events Involving the Botulinum Neurotoxin Complex Genes in Clostridium botulinum Types A, B, E and F and Clostridium butyricum Type E Strains. BMC Biol. 2009, 7, 66. [Google Scholar] [CrossRef]
  111. Sharma, V.K.; Akavaram, S.; Schaut, R.G.; Bayles, D.O. Comparative Genomics Reveals Structural and Functional Features Specific to the Genome of a Foodborne Escherichia coli O157:H7. BMC Genom. 2019, 20, 196. [Google Scholar] [CrossRef]
  112. Losada, L.; Ronning, C.M.; Deshazer, D.; Woods, D.; Fedorova, N.; Kim, H.S.; Shabalina, S.A.; Pearson, T.R.; Brinkac, L.; Tan, P.; et al. Continuing Evolution of Burkholderia mallei through Genome Reduction and Large-Scale Rearrangements. Genome Biol. Evol. 2010, 2, 102–116. [Google Scholar] [CrossRef]
  113. Chen, Y.; Stine, O.C.; Badger, J.H.; Gil, A.I.; Nair, G.B.; Nishibuchi, M.; Fouts, D.E. Comparative Genomic Analysis of Vibrio parahaemolyticus: Serotype Conversion and Virulence. BMC Genom. 2011, 12, 294. [Google Scholar] [CrossRef] [PubMed]
  114. Comandatore, F.; Sassera, D.; Bayliss, S.C.; Scaltriti, E.; Gaiarsa, S.; Cao, X.; Gales, A.C.; Saito, R.; Pongolini, S.; Brisse, S.; et al. Gene Composition as a Potential Barrier to Large Recombinations in the Bacterial Pathogen Klebsiella pneumoniae. Genome Biol. Evol. 2019, 11, 3240–3251. [Google Scholar] [CrossRef] [PubMed]
  115. Gomez-Valero, L.; Rusniok, C.; Jarraud, S.; Vacherie, B.; Rouy, Z.; Barbe, V.; Medigue, C.; Etienne, J.; Buchrieser, C. Extensive Recombination Events and Horizontal Gene Transfer Shaped the Legionella pneumophila Genomes. BMC Genom. 2011, 12, 536. [Google Scholar] [CrossRef]
  116. Campisi, E.; Rinaudo, C.D.; Donati, C.; Barucco, M.; Torricelli, G.; Edwards, M.S.; Baker, C.J.; Margarit, I.; Rosini, R. Serotype IV Streptococcus agalactiae ST-452 Has Arisen from Large Genomic Recombination Events between CC23 and the Hypervirulent CC17 Lineages. Sci. Rep. 2016, 6, 29799. [Google Scholar] [CrossRef] [PubMed]
  117. Gawlik, D.; Ruppelt-Lorz, A.; Müller, E.; Reißig, A.; Hotzel, H.; Braun, S.D.; Söderquist, B.; Ziegler-Cordts, A.; Stein, C.; Pletz, M.W.; et al. Molecular Investigations on a Chimeric Strain of Staphylococcus aureus Sequence Type 80. PLoS ONE 2020, 15, e0232071. [Google Scholar] [CrossRef]
  118. Spoor, L.E.; Richardson, E.; Richards, A.C.; Wilson, G.J.; Mendonca, C.; Gupta, R.K.; McAdam, P.R.; Nutbeam-Tuffs, S.; Black, N.S.; O’gara, J.P.; et al. Recombination-Mediated Remodelling of Host–Pathogen Interactions during Staphylococcus aureus Niche Adaptation. Microb. Genom. 2015, 1, e36. [Google Scholar] [CrossRef]
  119. Lista, F.; Faggioni, G.; Valjevac, S.; Ciammaruconi, A.; Vaissaire, J.; Le Doujet, C.; Gorgé, O.; De Santis, R.; Carattoli, A.; Ciervo, A.; et al. Genotyping of Bacillus anthracis Strains Based on Automated Capillary 25-Loci Multiple Locus Variable-Number Tandem Repeats Analysis. BMC Microbiol. 2006, 6, 33. [Google Scholar] [CrossRef]
  120. Spuesens, E.B.M.; Oduber, M.; Hoogenboezem, T.; Stuijter, M.; Hartwig, N.G.; van Rossum, A.M.C.; Vink, C. Sequence Variations in RepMP2/3 and RepMP4 Elements Reveal Intragenomic Homologous DNA Recombination Events in Mycoplasma pneumoniae. Microbiology 2009, 155, 2182–2196. [Google Scholar] [CrossRef]
  121. Partridge, S.R.; Kwong, S.M.; Firth, N.; Jensen, S.O. Mobile Genetic Elements Associated with Antimicrobial Resistance. Clin. Microbiol. Rev. 2018, 31, e00088-17. [Google Scholar] [CrossRef]
  122. Croucher, N.J.; Hanage, W.P.; Harris, S.R.; McGee, L.; van der Linden, M.; de Lencastre, H.; Sá-Leão, R.; Song, J.-H.; Ko, K.S.; Beall, B.; et al. Variable Recombination Dynamics during the Emergence, Transmission and “disarming” of a Multidrug-Resistant Pneumococcal Clone. BMC Biol. 2014, 12, 49. [Google Scholar] [CrossRef]
  123. Rolo, J.; Worning, P.; Nielsen, J.B.; Bowden, R.; Bouchami, O.; Damborg, P.; Guardabassi, L.; Perreten, V.; Tomasz, A.; Westh, H.; et al. Evolutionary Origin of the Staphylococcal Cassette Chromosome Mec (SCC Mec). Antimicrob. Agents Chemother. 2017, 61, e02302-16. [Google Scholar] [CrossRef] [PubMed]
  124. Araki, H.; Innan, H.; Kreitman, M.; Bergelson, J. Molecular Evolution of Pathogenicity-Island Genes in Pseudomonas viridiflava. Genetics 2007, 177, 1031–1041. [Google Scholar] [CrossRef] [PubMed]
  125. Li, M.; Shen, X.; Yan, J.; Han, H.; Zheng, B.; Liu, D.; Cheng, H.; Zhao, Y.; Rao, X.; Wang, C.; et al. GI-Type T4SS-Mediated Horizontal Transfer of the 89K Pathogenicity Island in Epidemic Streptococcus suis Serotype 2. Mol. Microbiol. 2011, 79, 1670–1683. [Google Scholar] [CrossRef]
  126. Elliott, B.; Dingle, K.E.; Didelot, X.; Crook, D.W.; Riley, T.V. The Complexity and Diversity of the Pathogenicity Locus in Clostridium difficile Clade 5. Genome Biol. Evol. 2014, 6, 3159–3170. [Google Scholar] [CrossRef] [PubMed]
  127. Planet, P.J.; Kachlany, S.C.; Fine, D.H.; DeSalle, R.; Figurski, D.H. The Widespread Colonization Island of Actinobacillus actinomycetemcomitans. Nat. Genet. 2003, 34, 193–198. [Google Scholar] [CrossRef]
  128. Schubert, S.; Nörenberg, D.; Clermont, O.; Magistro, G.; Wieser, A.; Romann, E.; Hoffmann, C.; Weinert, K.; Denamur, E. Prevalence and Phylogenetic History of the TcpC Virulence Determinant in Escherichia coli. Int. J. Med. Microbiol. 2010, 300, 429–434. [Google Scholar] [CrossRef]
  129. Thrane, S.W.; Taylor, V.L.; Freschi, L.; Kukavica-Ibrulj, I.; Boyle, B.; Laroche, J.; Pirnay, J.P.; Lévesque, R.C.; Lam, J.S.; Jelsbaka, L. The Widespread Multidrug-Resistant Serotype O12 Pseudomonas aeruginosa Clone Emerged through Concomitant Horizontal Transfer of Serotype Antigen and Antibiotic Resistance Gene Clusters. mBio 2015, 6, e01396-15. [Google Scholar] [CrossRef]
  130. Liao, J.; Orsi, R.H.; Carroll, L.M.; Kovac, J.; Ou, H.; Zhang, H.; Wiedmann, M. Serotype-Specific Evolutionary Patterns of Antimicrobial-Resistant Salmonella enterica. BMC Evol. Biol. 2019, 19, 132. [Google Scholar] [CrossRef]
  131. Morales, M.; García, P.; De La Campa, A.G.; LinÍares, J.; Ardanuy, C.; GarciÍa, E. Evidence of Localized Prophage-Host Recombination in the lytA Gene, Encoding the Major Pneumococcal Autolysin. J. Bacteriol. 2010, 192, 2624–2632. [Google Scholar] [CrossRef]
  132. Tamarit, D.; Neuvonen, M.M.; Engel, P.; Guy, L.; Andersson, S.G.E. Origin and Evolution of the Bartonella Gene Transfer Agent. Mol. Biol. Evol. 2018, 35, 451–464. [Google Scholar] [CrossRef]
  133. Verne, S.; Johnson, M.; Bouchon, D.; Grandjean, F. Evidence for Recombination between Feminizing Wolbachia in the Isopod Genus Armadillidium. Gene 2007, 397, 58–66. [Google Scholar] [CrossRef]
  134. Miller, J.C.; Stevenson, B. Immunological and Genetic Characterization of Borrelia burgdorferi BapA and EppA Proteins. Microbiology 2003, 149, 1113–1125. [Google Scholar] [CrossRef]
  135. Akins, D.R.; Caimano, M.J.; Yang, X.; Cerna, F.; Norgard, M.V.; Radolf, J.D. Molecular and Evolutionary Analysis of Borrelia burgdorferi 297 Circular Plasmid-Encoded Lipoproteins with OspE- and OspF-like Leader Peptides. Infect. Immun. 1999, 67, 1526–1532. [Google Scholar] [CrossRef]
  136. Joseph, S.J.; Didelot, X.; Gandhi, K.; Dean, D.; Read, T.D. Interplay of Recombination and Selection in the Genomes of Chlamydia trachomatis. Biol. Direct 2011, 6, 28. [Google Scholar] [CrossRef] [PubMed]
  137. Gomes, J.P.; Bruno, W.J.; Nunes, A.; Santos, N.; Florindo, C.; Borrego, M.J.; Dean, D. Evolution of Chlamydia trachomatis Diversity Occurs by Widespread Interstrain Recombination Involving Hotspots. Genome Res. 2007, 17, 50–60. [Google Scholar] [CrossRef] [PubMed]
  138. Smelov, V.; Vrbanac, A.; van Ess, E.F.; Noz, M.P.; Wan, R.; Eklund, C.; Morgan, T.; Shrier, L.A.; Sanders, B.; Dillner, J.; et al. Chlamydia trachomatis Strain Types Have Diversified Regionally and Globally with Evidence for Recombination across Geographic Divides. Front. Microbiol. 2017, 8, 2195. [Google Scholar] [CrossRef]
  139. Roulis, E.; Bachmann, N.; Humphrys, M.; Myers, G.; Huston, W.; Polkinghorne, A.; Timms, P. Phylogenetic Analysis of Human Chlamydia pneumoniae Strains Reveals a Distinct Australian Indigenous Clade That Predates European Exploration of the Continent. BMC Genom. 2015, 16, 1094. [Google Scholar] [CrossRef] [PubMed]
  140. Mendoza-Elizalde, S.; Cortés-Márquez, A.C.; Zuñiga, G.; Cerritos, R.; Valencia-Mayoral, P.; Sánchez, A.C.; Olivares-Clavijo, H.; Velázquez-Guadarrama, N. Inference from the Analysis of Genetic Structure of Helicobacter pylori Strains Isolates from Two Paediatric Patients with Recurrent Infection. BMC Microbiol. 2019, 19, 184. [Google Scholar] [CrossRef]
  141. Tsai, Y.H.L.; Orsi, R.H.; Nightingale, K.K.; Wiedmann, M. Listeria monocytogenes Internalins Are Highly Diverse and Evolved by Recombination and Positive Selection. Infect. Genet. Evol. 2006, 6, 378–389. [Google Scholar] [CrossRef]
  142. Hill, D.J.; Whittles, C.; Virji, M. A Novel Group of Moraxella catarrhalis UspA Proteins Mediates Cellular Adhesion via CEACAMs and Vitronectin. PLoS ONE 2012, 7, e45452. [Google Scholar] [CrossRef]
  143. Muzzi, A.; Moschioni, M.; Covacci, A.; Rappuoli, R.; Donati, C. Pilus Operon Evolution in Streptococcus pneumoniae Is Driven by Positive Selection and Recombination. PLoS ONE 2008, 3, e3660. [Google Scholar] [CrossRef] [PubMed]
  144. Pandya, G.A.; McEllistrem, M.C.; Venepally, P.; Holmes, M.H.; Jarrahi, B.; Sanka, R.; Liu, J.; Karamycheva, S.A.; Bai, Y.; Fleischmann, R.D.; et al. Monitoring the Long-Term Molecular Epidemiology of the Pneumococcus and Detection of Potential “vaccine Escape” Strains. PLoS ONE 2011, 6, e15950. [Google Scholar] [CrossRef] [PubMed]
  145. Rahbar, M.R.; Zarei, M.; Jahangiri, A.; Khalili, S.; Nezafat, N.; Negahdaripour, M.; Fattahian, Y.; Ghasemi, Y. Trimeric Autotransporter Adhesins in Acinetobacter baumannii, Coincidental Evolution at Work. Infect. Genet. Evol. 2019, 71, 116–127. [Google Scholar] [CrossRef] [PubMed]
  146. Achtman, M. Clonal Spread of Serogroup A Meningococci: A Paradigm for the Analysis of Microevolution in Bacteria. Mol. Microbiol. 1994, 11, 15–22. [Google Scholar] [CrossRef] [PubMed]
  147. David, S.; Sánchez-Busó, L.; Harris, S.R.; Marttinen, P.; Rusniok, C.; Buchrieser, C.; Harrison, T.G.; Parkhill, J. Dynamics and Impact of Homologous Recombination on the Evolution of Legionella pneumophila. PLoS Genet. 2017, 13, e1006855. [Google Scholar] [CrossRef]
  148. Tsuru, T.; Kobayashi, I. Multiple Genome Comparison within a Bacterial Species Reveals a Unit of Evolution Spanning Two Adjacent Genes in a Tandem Paralog Cluster. Mol. Biol. Evol. 2008, 25, 2457–2473. [Google Scholar] [CrossRef]
  149. Brochet, M.; Couvé, E.; Zouine, M.; Vallaeys, T.; Rusniok, C.; Lamy, M.C.; Buchrieser, C.; Trieu-Cuot, P.; Kunst, F.; Poyart, C.; et al. Genomic Diversity and Evolution within the Species Streptococcus agalactiae. Microbes Infect. 2006, 8, 1227–1243. [Google Scholar] [CrossRef]
  150. Whatmore, A.M.; Kapur, V.; Sullivan, D.J.; Musser, J.M.; Kehoe, M.A. Non-congruent Relationships between Variation in Emm Gene Sequences and the Population Genetic Structure of Group A Streptococci. Mol. Microbiol. 1994, 14, 619–631. [Google Scholar] [CrossRef]
  151. McBride, A.J.A.; Cerqueira, G.M.; Suchard, M.A.; Moreira, A.N.; Zuerner, R.L.; Reis, M.G.; Haake, D.A.; Ko, A.I.; Dellagostin, O.A. Genetic Diversity of the Leptospiral Immunoglobulin-like (Lig) Genes in Pathogenic Leptospira spp. Infect. Genet. Evol. 2009, 9, 196–205. [Google Scholar] [CrossRef]
  152. Sabat, A.J.; Wladyka, B.; Kosowska-Shick, K.; Grundmann, H.; Van Dijl, J.M.; Kowal, J.; Appelbaum, P.C.; Dubin, A.; Hryniewicz, W. Polymorphism, Genetic Exchange and Intragenic Recombination of the Aureolysin Gene among Staphylococcus aureus Strains. BMC Microbiol. 2008, 8, 129. [Google Scholar] [CrossRef]
  153. Watanabe, S.; Ito, T.; Sasaki, T.; Li, S.; Uchiyama, I.; Kishii, K.; Kikuchi, K.; Skov, R.L.; Hiramatsu, K. Genetic Diversity of Staphylocoagulase Genes (coa): Insight into the Evolution of Variable Chromosomal Virulence Factors in Staphylococcus aureus. PLoS ONE 2009, 4, e5714. [Google Scholar] [CrossRef]
  154. Muzzi, A.; Mora, M.; Pizza, M.; Rappuoli, R.; Donati, C. Conservation of Meningococcal Antigens in the Genus Neisseria. mBio 2013, 4, e00163-13. [Google Scholar] [CrossRef]
  155. Gray, R.R.; Mulligan, C.J.; Molini, B.J.; Sun, E.S.; Giacani, L.; Godornes, C.; Kitchen, A.; Lukehart, S.A.; Centurion-Lara, A. Molecular Evolution of the tprC, D, I, K, G, and J Genes in the Pathogenic Genus Treponema. Mol. Biol. Evol. 2006, 23, 2220–2233. [Google Scholar] [CrossRef]
  156. Guttman, D.S.; Gropp, S.J.; Morgan, R.L.; Wang, P.W. Diversifying Selection Drives the Evolution of the Type III Secretion System Pilus of Pseudomonas syringae. Mol. Biol. Evol. 2006, 23, 2342–2354. [Google Scholar] [CrossRef] [PubMed]
  157. Costa, J.; Teixeira, P.G.; D’Avó, A.F.; Júnior, C.S.; Veríssimo, A. Intragenic Recombination Has a Critical Role on the Evolution of Legionella pneumophila Virulence-Related Effector SidJ. PLoS ONE 2014, 9, e109840. [Google Scholar] [CrossRef] [PubMed]
  158. Costa, J.; Tiago, I.; da Costa, M.S.; Veríssimo, A. Molecular Evolution of Legionella pneumophila DotA Gene, the Contribution of Natural Environmental Strains. Environ. Microbiol. 2010, 12, 2711–2729. [Google Scholar] [CrossRef] [PubMed]
  159. Lee, I.; Davies, R.L. Evidence for a Common Gene Pool and Frequent Recombinational Exchange of the tbpBA Operon in Mannheimia haemolytica, Mannheimia glucosida and Bibersteinia trehalosi. Microbiology 2011, 157, 123–135. [Google Scholar] [CrossRef] [PubMed]
  160. Harrison, O.B.; Maiden, M.C.J.; Rokbi, B. Distribution of Transferrin Binding Protein B Gene (tbpB) Variants among Neisseria Species. BMC Microbiol. 2008, 8, 66. [Google Scholar] [CrossRef] [PubMed]
  161. Giffard, P.M.; Allen, D.M.; Milward, C.P.; Simpson, C.L.; Jacques, N.A. Sequence of the GtfK Gene of Streptococcus salivarius ATCC 25975 and Evolution of the gtf Genes of Oral Streptococci. J. Gen. Microbiol. 1993, 139, 1511–1522. [Google Scholar] [CrossRef]
  162. Ng, V.; Lin, W.J. Comparison of Assembled Clostridium botulinum A1 Genomes Revealed Their Evolutionary Relationship. Genomics 2014, 103, 94–106. [Google Scholar] [CrossRef]
  163. Prisilla, A.; Prathiviraj, R.; Chellapandi, P. Molecular Evolutionary Constraints That Determine the Avirulence State of Clostridium botulinum C2 Toxin. J. Mol. Evol. 2017, 84, 174–186. [Google Scholar] [CrossRef] [PubMed]
  164. Mansfield, M.J.; Tremblay, B.J.M.; Zeng, J.; Wei, X.; Hodgins, H.; Worley, J.; Bry, L.; Dong, M.; Doxey, A.C. Phylogenomics of 8,839 Clostridioides difficile Genomes Reveals Recombination-Driven Evolution and Diversification of Toxin A and B. PLoS Pathog. 2020, 16, e1009181. [Google Scholar] [CrossRef]
  165. Davies, R.L.; Campbell, S.; Whittam, T.S. Mosaic Structure and Molecular Evolution of the Leukotoxin Operon (lktCABD) in Mannheimia (Pasteurella) haemolytica, Mannheimia glucosida, and Pasteurella trehalosi. J. Bacteriol. 2002, 184, 266–277. [Google Scholar] [CrossRef]
  166. Turner, C.E.; Holden, M.T.G.; Blane, B.; Horner, C.; Peacock, S.J.; Sriskandan, S. The Emergence of Successful Streptococcus pyogenes Lineages through Convergent Pathways of Capsule Loss and Recombination Directing High Toxin Expression. mBio 2019, 10, e02521-19. [Google Scholar] [CrossRef] [PubMed]
  167. Jobling, M.G.; Holmes, R.K. Type II Heat-Labile Enterotoxins from 50 Diverse Escherichia coli Isolates Belong Almost Exclusively to the LT-IIc Family and May Be Prophage Encoded. PLoS ONE 2012, 7, e29898. [Google Scholar] [CrossRef] [PubMed]
  168. Duncan, S.S.; Valk, P.L.; Shaffer, C.L.; Bordenstein, S.R.; Cover, T.L. J-Western Forms of Helicobacter pylori cagA Constitute a Distinct Phylogenetic Group with a Widespread Geographic Distribution. J. Bacteriol. 2012, 194, 1593–1604. [Google Scholar] [CrossRef]
  169. Bonnin, R.A.; Poirel, L.; Nordmann, P. New Delhi Metallo-β-Lactamase-Producing Acinetobacter baumannii: A Novel Paradigm for Spreading Antibiotic Resistance Genes. Future Microbiol. 2014, 9, 33–41. [Google Scholar] [CrossRef]
  170. Liu, L.; Cui, Y.; Zheng, B.; Jiang, S.; Yu, W.; Shen, P.; Ji, J.; Li, L.; Qin, N.; Xiao, Y. Analysis of Tigecycline Resistance Development in Clinical Acinetobacter baumannii Isolates through a Combined Genomic and Transcriptomic Approach. Sci. Rep. 2016, 6, 26930. [Google Scholar] [CrossRef]
  171. Diaz Caballero, J.; Clark, S.T.; Wang, P.W.; Donaldson, S.L.; Coburn, B.; Tullis, D.E.; Yau, Y.C.W.; Waters, V.J.; Hwang, D.M.; Guttman, D.S. A Genome-Wide Association Analysis Reveals a Potential Role for Recombination in the Evolution of Antimicrobial Resistance in Burkholderia multivorans. PLoS Pathog. 2018, 14, e1007453. [Google Scholar] [CrossRef]
  172. Tchesnokova, V.; Radey, M.; Chattopadhyay, S.; Larson, L.; Weaver, J.L.; Kisiela, D.; Sokurenko, E.V. Pandemic Fluoroquinolone Resistant Escherichia coli Clone ST1193 Emerged via Simultaneous Homologous Recombinations in 11 Gene Loci. Proc. Natl. Acad. Sci. USA 2019, 116, 14740–14748. [Google Scholar] [CrossRef]
  173. De Been, M.; Van Schaik, W.; Cheng, L.; Corander, J.; Willems, R.J. Recent Recombination Events in the Core Genome Are Associated with Adaptive Evolution in Enterococcus faecium. Genome Biol. Evol. 2013, 5, 1524–1535. [Google Scholar] [CrossRef] [PubMed]
  174. Sugiyama, T.; Kido, N.; Kato, Y.; Koide, N.; Yoshida, T.; Yokochi, T. Evolutionary Relationship among rfb Gene Clusters Synthesizing Mannose Homopolymer as O-Specific Polysaccharides in Escherichia coli and Klebsiella. Gene 1997, 198, 111–113. [Google Scholar] [CrossRef] [PubMed]
  175. Yahara, K.; Kawai, M.; Furuta, Y.; Takahashi, N.; Handa, N.; Tsuru, T.; Oshima, K.; Yoshida, M.; Azuma, T.; Hattori, M.; et al. Genome-Wide Survey of Mutual Homologous Recombination in a Highly Sexual Bacterial Species. Genome Biol. Evol. 2012, 4, 628–640. [Google Scholar] [CrossRef] [PubMed]
  176. Holt, K.; Kenyon, J.J.; Hamidian, M.; Schultz, M.B.; Pickard, D.J.; Dougan, G.; Hall, R. Five Decades of Genome Evolution in the Globally Distributed, Extensively Antibiotic-Resistant Acinetobacter baumannii Global Clone 1. Microb. Genom. 2016, 2, e000052. [Google Scholar] [CrossRef]
  177. Rishishwar, L.; Katz, L.S.; Sharma, N.V.; Rowe, L.; Frace, M.; Thomas, J.D.; Harcourt, B.H.; Mayer, L.W.; Jordana, I.K. Genomic Basis of a Polyagglutinating Isolate of Neisseria meningitidis. J. Bacteriol. 2012, 194, 5649–5656. [Google Scholar] [CrossRef]
  178. Berti, F.; Campisi, E.; Toniolo, C.; Morelli, L.; Crotti, S.; Rosini, R.; Romano, M.R.; Pinto, V.; Brogioni, B.; Torricelli, G.; et al. Structure of the Type IX Group B Streptococcus Capsular Polysaccharide and Its Evolutionary Relationship with Types V and VII. J. Biol. Chem. 2014, 289, 23437–23448. [Google Scholar] [CrossRef]
  179. Neemuchwala, A.; Teatero, S.; Athey, T.B.T.; McGeer, A.; Fittipaldi, N. Capsular Switching and Other Large-Scale Recombination Events in Invasive Sequence Type 1 Group B Streptococcus. Emerg. Infect. Dis. 2016, 22, 1941–1944. [Google Scholar] [CrossRef]
  180. Croucher, N.J.; Kagedan, L.; Thompson, C.M.; Parkhill, J.; Bentley, S.D.; Finkelstein, J.A.; Lipsitch, M.; Hanage, W.P. Selective and Genetic Constraints on Pneumococcal Serotype Switching. PLoS Genet. 2015, 11, e1005095. [Google Scholar] [CrossRef]
  181. Alqasim, A.; Scheutz, F.; Zong, Z.; McNally, A. Comparative Genome Analysis Identifies Few Traits Unique to the Escherichia coli ST131 H30Rx Clade and Extensive Mosaicism at the Capsule Locus. BMC Genom. 2014, 15, 830. [Google Scholar] [CrossRef]
  182. Starcevic, A.; Diminic, J.; Zucko, J.; Elbekali, M.; Schlosser, T.; Lisfi, M.; Vukelic, A.; Long, P.F.; Hranueli, D.; Cullum, J. A Novel Docking Domain Interface Model Predicting Recombination between Homoeologous Modular Biosynthetic Gene Clusters. J. Ind. Microbiol. Biotechnol. 2011, 38, 1295–1304. [Google Scholar] [CrossRef]
Figure 1. Functional impact of loci subjected to recombination and their genomic context. In each case, the color corresponds to the consequences of recombination events, classified into three main groups, namely, ecological adaptation, symbiosis, and pathogenesis. The list of the analyzed events from 91 bacterial species is presented in Table S1. Used abbreviations are as follows: ICEs—integrative and conjugative elements; ISs—insertion sequences; Gis—genomic islands. (a) The absolute group-wise number of events attributed to different types of genomic regions. Regular loci imply genes and/or operons that are not surrounded by or located within mobile genetic elements. (b) The distribution of loci types encoding protein products attributed to their functional roles. Each tile implies that a certain protein group is encoded by a particular genomic locus. (c) A schematic pictorial representation of the structures and processes in which proteins encoded by loci subjected to recombination take part.
Figure 1. Functional impact of loci subjected to recombination and their genomic context. In each case, the color corresponds to the consequences of recombination events, classified into three main groups, namely, ecological adaptation, symbiosis, and pathogenesis. The list of the analyzed events from 91 bacterial species is presented in Table S1. Used abbreviations are as follows: ICEs—integrative and conjugative elements; ISs—insertion sequences; Gis—genomic islands. (a) The absolute group-wise number of events attributed to different types of genomic regions. Regular loci imply genes and/or operons that are not surrounded by or located within mobile genetic elements. (b) The distribution of loci types encoding protein products attributed to their functional roles. Each tile implies that a certain protein group is encoded by a particular genomic locus. (c) A schematic pictorial representation of the structures and processes in which proteins encoded by loci subjected to recombination take part.
Toxins 15 00568 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shikov, A.E.; Savina, I.A.; Nizhnikov, A.A.; Antonets, K.S. Recombination in Bacterial Genomes: Evolutionary Trends. Toxins 2023, 15, 568. https://doi.org/10.3390/toxins15090568

AMA Style

Shikov AE, Savina IA, Nizhnikov AA, Antonets KS. Recombination in Bacterial Genomes: Evolutionary Trends. Toxins. 2023; 15(9):568. https://doi.org/10.3390/toxins15090568

Chicago/Turabian Style

Shikov, Anton E., Iuliia A. Savina, Anton A. Nizhnikov, and Kirill S. Antonets. 2023. "Recombination in Bacterial Genomes: Evolutionary Trends" Toxins 15, no. 9: 568. https://doi.org/10.3390/toxins15090568

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop