Skip to main content

MINI REVIEW article

Front. Microbiol., 31 August 2018
Sec. Infectious Agents and Disease
This article is part of the Research Topic The Deadly Secrets of C. Difficile - Insights into Host-Pathogen Interaction View all 28 articles

Bacteriophages Contribute to Shaping Clostridioides (Clostridium) difficile Species

  • Department of Microbiology and Infectious Diseases, Faculty of Medicine and Health Sciences, Université de Sherbrooke, Sherbrooke, QC, Canada

Bacteriophages (phages) are bacterial viruses that parasitize bacteria. They are highly prevalent in nature, with an estimated 1031 viral particles in the whole biosphere, and they outnumber bacteria by at least 10-fold. Hence, phages represent important drivers of bacterial evolution, although our knowledge of the role played by phages in the mammalian gut is still embryonic. Several pathogens owe their virulence to the integrated phages (prophages) they harbor, which encode diverse virulence factors such as toxins. Clostridioides (Clostridium) difficile is an important opportunistic pathogen and several phages infecting this species have been described over the last decade. However, their exact contribution to the biology and virulence of this pathogen remains elusive. Current data have shown that C. difficile phages can alter virulence-associated phenotypes, in particular toxin production, by interfering with bacterial regulatory circuits through crosstalk with phage proteins for example. One phage has also been found to encode a complete binary toxin locus. Multiple regulatory genes have also been identified in phage genomes, suggesting that their impact on the host can be complex and often subtle. In this minireview, the current state of knowledge, major findings, and pending questions regarding C. difficile phages will be presented. In addition, with the apparent role played by phages in the success of fecal microbiota transplantation and the perspective of phage therapy for treatment of recurrent C. difficile infection, it has become even more crucial to understand what C. difficile phages do in the gut, how they impact their host, and how they influence the epidemiology and evolution of this clinically important pathogen.

Bacteriophages

Bacteriophages (phages) are bacterial viruses that infect bacteria. At ∼1031 viral particles, they represent the most abundant biological entities in the biosphere and almost all bacteria are susceptible to phage attacks. Phages are therefore important drivers of bacterial evolution (Brüssow and Hendrix, 2002). Depending on the phages’ replication strategy, their impact on bacterial populations can be drastically different. The two most frequent mechanisms of phage replication are the lytic cycle, and the lysogenic cycle (Figure 1). Phages that replicate only via the lytic pathway are referred to as “virulent” and inevitably lead to death of the host upon infection. Those phages that can replicate either by the lytic or the lysogenic cycle are said to be “temperate.” When temperate phages become integrated into the host genome (i.e., prophages) their host is said to be lysogenic.

FIGURE 1
www.frontiersin.org

FIGURE 1. Summary of the replication strategies and multiple impacts of phages infecting C. difficile.

Bacteriophages Infecting C. Difficile

Most phages infecting Clostridioides (Clostridium) difficile were isolated following induction of prophages (Shan et al., 2012; Hargreaves and Clokie, 2014; Sekulovic et al., 2014). However, free phages have also been isolated directly from fecal supernatants of patients infected with C. difficile, hence prophage induction occurs in vivo (Meessen-Pinard et al., 2012). In the 1980s and 1990s, C. difficile phages were studied for their potential as strain typing tools (Sell et al., 1983; Dei, 1989; Mahony et al., 1991). Now their potential for phage therapy applications is being explored (Nale et al., 2016b, 2018). At the time of writing this manuscript, at least 24 complete phage genomes were available in public databases (Table 1). Most of them are members of the Myoviridae family of the order Caudovirales (i.e., phages with contractile tails), and six phages are members of the Siphoviridae family (i.e., phages with long non-contractile tails) (Ackermann and Prangishvili, 2012). Functional data describing the lytic cycle of C. difficile phages are quite scarce. The few one-step growth curve experiments published so far suggest highly variable latency periods (from ∼30 min to 2 h), (Goh et al., 2005b; Sekulovic et al., 2011) and burst sizes (i.e., virions released per infected cell), with as few as 5 phages/cell for phage ϕC2 (Goh et al., 2005b), to 122 phages/cell for phage 56 (Mahony et al., 1985). In addition, most phages have relatively narrow host ranges (Goh et al., 2005b; Sekulovic et al., 2011, 2014; Rashid et al., 2016), which is directly related to the availability of a suitable host receptor, the presence of endogenous prophages conferring resistance, and the presence of antiphage systems including clustered regularly interspaced short palindromic repeat sequences (CRISPRs) (Hargreaves et al., 2014a; Boudry et al., 2015), the recently described superinfection exclusion system CwpV (Sekulovic et al., 2015), and possibly others. Of particular interest, the receptor(s) used by C. difficile phages to infect their host remain(s) to be clearly identified. In other Gram-positive bacteria, different cell surface components are used as phage receptors, from single proteins to polysaccharides or teichoic acids. As examples, the Bacillus subtilis YueB (São-José et al., 2006) and Lactococcus lactis Pip (Babu et al., 1995) proteins are, respectively, used by phages SPP1 and c2 to infect their host. Different polysaccharides composing the pellicle are used as receptors by lactococcal phages such as p2 (Bebeacua et al., 2013). Of note, recent data with Diffocins, i.e., phage tail-like bacteriocins that resemble Myoviridae phage tails and that kill their host by puncturing the cell, point to a central role of the surface layer protein A (SlpA) as a general phage receptor used by Diffocins and Myoviridae phages (Gebhart et al., 2015; Kirk et al., 2017). Detailed molecular interactions between phages and the C. difficile surface need to be further investigated, in particular regarding Siphoviridae phages, and considering the potential of phages as therapeutic agents.

TABLE 1
www.frontiersin.org

TABLE 1. List and relevant characteristics of C. difficile phages for which a complete genome sequence is available in GenBank.

High Prevalence of Prophages in C. Difficile Genomes

Over 1,300 C. difficile genomes have been fully sequenced and are available in public repositories, but thousands of additional genomes have also been sequenced and are available through collaborative research (Garneau et al., 2018). Although uncommon, as many as 5–6 different prophages were identified in a single C. difficile genome (Amy et al., 2018; Ramírez-Vargas et al., 2018). However, between 1 and 3 prophages are more frequently observed, in addition to genomic “islands” containing phage-related genes. Recent studies highlighted the prevalence of large phage genomes that reside as extrachromosomal DNA in C. difficile (Garneau et al., 2018; Ramírez-Vargas et al., 2018). For example, the large phiCD211/phiCDIF1297T and related phages, with genomes of ≥131-kb, have been detected in 5% of 2,584 C. difficile genomes analyzed, spanning 21 different multi-locus sequence types (MLST) (Wittmann et al., 2015; Garneau et al., 2018). Ten other large phage genomes (∼128–135-kb), including phiCD5763, phiCD5774, and phiCD2955, were recently described in C. difficile isolates from around the world and representing seven different MLST sequence types (Ramírez-Vargas et al., 2018). Comparative genomic analyses underlined the important genetic variability among large phages, and they could eventually be used as genetic markers to subtype and monitor specific strains during epidemiological studies, as suggested for Salmonella enterica (Mottawea et al., 2018).

It is worth mentioning that extrachromosomal phage genomes can be difficult to differentiate from large plasmids containing phage genes. A study by Amy et al. (2018) reported the characterization of a large plasmid in C. difficile strain DLL3026. This 46-kb plasmid, called pDLL3026, and several other plasmids of similar size identified in other isolates, harbor a significant number of phage structural genes coding for head and tail morphogenesis, recombinases/integrases and phage regulators. The presence of partition genes like parM and parR and DNA similarity with plasmids led the authors to conclude that these were plasmids. However, the presence of partition genes like parA has also been reported in other phages, including phiCD6356 (Horgan et al., 2010), ϕCD38-2 (Sekulovic et al., 2011), and phiSemix9P1 (Riedel et al., 2017), the latter two known to be maintained as extrachromosomal DNAs in lysogenic cells. Large phage genomes such as phiCD211/phiCDIF1296T and phiCD5763 also seem to be frequently found as extrachromosomal DNA, and ParM homologs were identified in some of them (Garneau et al., 2018; Ramírez-Vargas et al., 2018). Therefore, in the absence of functional data to assess the inducibility and production of infectious particles from these large “plasmids,” it is hard to conclude on their exact nature.

The identification of complete prophages in bacterial genomes has been greatly improved, thanks to the development of tools such as PHAST and PHASTER (Arndt et al., 2016, 2017). But the task is more challenging with decaying prophage remnants that have lost many of the conserved phage components such as structural genes. Yet, these remnants could still influence their host even if they can’t replicate or produce complete infectious particles. Diffocins are a good example: these phage tail-like particles resemble Myoviridae phage tails, but lack a capsid and genetic material (Gebhart et al., 2012). They kill their host following induction and lysis of the cell, and also kill other competing cells around, but they can’t produce infectious particles. The functional role of Diffocins remains to be clarified, but they possibly provide a competitive advantage to C. difficile strains carrying them by killing surrounding competitors (Kirk et al., 2017).

The Consequences of Prophage Induction

The role of prophages in the physiology and virulence of C. difficile is a topic of great interest, considering their prevalence and diversity, and the historical role prophages played in the virulence of other bacterial pathogens (Brüssow et al., 2004; Fortier and Sekulovic, 2013). Prophage stability is critical because of the direct consequence on the viability of the host itself and susceptible surrounding strains/species that can be re-infected. Induction can occur spontaneously, but is promoted by common antibiotics and various environmental stresses (Rokney et al., 2008; Meessen-Pinard et al., 2012; Shan et al., 2012). Of note, prophage induction triggered by antibiotics promote horizontal gene transfer and spreading of antibiotic resistance genes in mice (Modi et al., 2013). In vitro, phage ϕC2 was shown to mediate transduction of the Tn6215 transposon between C. difficile strains (Goh et al., 2013). Differences in abundance and diversity of gut phages has also been associated with diseases and could be the result of prophage induction (Norman et al., 2015; Manrique et al., 2016). Hence, better understanding the role of prophage induction in complex ecosystems such as the gut is of great interest.

Induction of prophages and phage-related elements can also have other important physiological roles in C. difficile. For example, excision of the phage-related mobile element called skinCd is important during the sporulation process (Haraldsen and Sonenshein, 2003; Saujet et al., 2014). The skinCd is a putative prophage remnant similar to the one identified in B. subtilis (skinBs) that interrupts the coding sequence of SigK, a sporulation-associated alternative sigma factor. Excision of the skinCd element at a specific time point during the sporulation process restores the coding sequence of the gene, allowing expression of sigK (Saujet et al., 2014; Fimlaid and Shen, 2015). Control of the excision of the skin element remains unclear, but a putative site-specific recombinase similar to SpoIVCA, encoded by cd1231 and located within skinCd, is suspected to be involved. Some crosstalk between prophages has been reported (Lemire et al., 2011; Sekulovic and Fortier, 2015), including between their recombinases (Singh et al., 2014). Therefore, other phage-encoded recombinases could possibly participate in skinCd excision as well, and thus influence sporulation (Saujet et al., 2014).

Prophages Influence Toxin Production in C. Difficile

The main virulence factors of C. difficile are the large TcdA and TcdB exotoxins. They are encoded on a 19.6-kb pathogenicity locus, the PaLoc (Rupnik et al., 2009). The PaLoc is thought to originate from an ancient prophage, since it shares a number of features with phages, in particular the tcdE gene encoding a phage-like holin involved in toxin secretion (Govind and Dupuy, 2012; Govind et al., 2015; Monot et al., 2015). Prophage induction per se has not been directly associated with toxin release or synthesis in C. difficile, as opposed to Shiga toxin-encoding phages in Escherichia coli (Kimmitt et al., 1999; Zhang et al., 2000) but some prophages interfere with toxin synthesis. For example, phage ϕCD119 was shown to express the RepR repressor, capable to bind a DNA region in the promoter of tcdR in the PaLoc, resulting in repression of toxin genes (Govind et al., 2009). On the contrary, phage ϕCD38-2 was shown to increase transcription of all five PaLoc genes by a yet unknown mechanism, resulting in more toxins produced in vitro. However, the impact of ϕCD38-2 on toxin synthesis was strain-dependent (Sekulovic et al., 2011) and similar observations were reported with other C. difficile phages (Goh et al., 2005a), suggesting that the influence of a prophage on its host partly depends on the genetic background.

A complete binary toxin locus (CdtLoc) has been recently identified in the genome of phage phiSemix9P1 (Riedel et al., 2017). The binary toxin, normally located on a 6.2-kb chromosomal locus called CdtLoc comprises 2 genes coding for the toxin components, cdtA and cdtB, as well as a regulator encoded by cdtR (Carman et al., 2011; Gerding et al., 2014). Of note, all three genes from the CdtLoc were shown to be transcribed from the phiSemix9P1 prophage, suggesting that it is functional, although no toxin assays have been performed (Riedel et al., 2017). The CdtLoc is present only in a subset of C. difficile isolates, including the epidemic ribotype 027 isolates (Bauer et al., 2011), and studies suggest that CDT contributes to virulence by promoting adhesion to epithelial cells (Schwan et al., 2009, 2014; Gerding et al., 2014). phiSemix9P1 has limited DNA homology with another C. difficile phage, ϕCD505, and the large pCDBI1 plasmid, suggesting that it is genetically unique. The identification of a CDT-encoding phage is intriguing but it might represent a rare isolated case, since it has never been observed in other C. difficile phages, including the numerous prophages identified in the course of genome sequencing projects. Nevertheless, it further supports the evolutionary role of phages in toxin conversion of C. difficile (Riedel et al., 2017).

Prophage Gene Expression During Lysogeny

During active phage replication, the transcriptional program of the host is profoundly restructured and metabolic resources are redirected toward phage replication. However, during the lysogenic cycle, prophages are generally quiescent and minimal gene transcription is observed from the prophage itself. Only a few gene products are required to establish and maintain lysogeny (Ainsworth et al., 2013), the CI repressor from the E. coli phage Lambda being the most well-characterized gene expressed during lysogeny (Oppenheim et al., 2005; Rokney et al., 2008).

Very little is known about transcriptional reprogramming during phage infection or lysogeny in C. difficile. In fact, only one study has looked at global gene expression during lysogeny (Sekulovic and Fortier, 2015). In that study, the ϕCD38-2 prophage was introduced into the epidemic strain R20291 and mRNA levels were assessed by RNAseq. It is important to mention that the prophage was maintained as a circular extrachromosomal DNA, so the host genome integrity was unaffected. On a genome-wide scale, the expression of 39 genes was significantly altered by the introduction of the prophage, including genes from the phi027 prophage already present in the host. This further supports the existence of some crosstalk between prophages (Sekulovic and Fortier, 2015). Two-thirds of the differentially expressed genes were downregulated twofold to threefold, and half of the differentially expressed genes were related to sugar uptake and metabolism, suggesting a possible impact on growth kinetics. Of note, the cwpV gene encoding a conserved surface protein was induced 20-fold in the lysogen. Transcription of cwpV is dependent on the configuration of a genetic switch located between the promoter and the gene. Recombination of the switch, catalyzed by the host-encoded RecV recombinase, turns transcription of cwpV ON or OFF in a phase-variable manner (Reynolds et al., 2011; Fagan and Fairweather, 2014). Only ∼5% of bacterial cells in culture express the CwpV protein at their surface, but in the R20291 lysogen carrying ϕCD38-2, this proportion increased to 95%, hence explaining the higher mRNA levels observed. The exact mechanism by which ϕCD38-2 influences phase variation remains unknown (Sekulovic and Fortier, 2015). CwpV is a large conserved cell wall protein suspected to contribute to cell adhesion and biofilm formation, and possibly immune evasion (Reynolds et al., 2011). The location of the protein at the cell surface and its apparent link with lysogeny suggested that it could play some role in phage infection. It turned out that CwpV has strong antiphage activity against several C. difficile phages of the Siphoviridae and Myoviridae families when overexpressed from a plasmid or from a “locked-ON” strain. Current data suggest that CwpV functions as a superinfection exclusion system (Sekulovic et al., 2015) that blocks phage DNA injection. The biological relevance of such an antiphage system seems obvious in the context of the gut microbiota. CwpV-ON strains would be protected from lytic phage attacks, which are expected to be relatively frequent in the gut due to high phage and bacterial densities (Manrique et al., 2016, 2017). Higher numbers of CwpV-ON cells could also contribute to colonization of the gut through increased bacterial adhesion and biofilm formation. Maybe of greater concern, however, is the fact that cwpV-expressing cells are naturally occurring in vitro due to phase variation and these cells are resistant to phage infection. Hence, looking at future phage therapy perspectives (Nale et al., 2016b), naturally occurring CwpV-positive cells in the gut could potentially compromise the efficacy of therapeutic phages. Further in vivo assays will be required to clarify the biological role and consequences of CwpV expression.

Other Impacts of Prophages on their Host

Several phage genomes carry cargo genes unrelated to the phage replication cycle, and their expression is often independent from the phage circuitry and occurs during lysogeny. The genes often code for virulence factors, including toxins, superantigens, and hydrolytic enzymes (Brüssow et al., 2004; Fortier and Sekulovic, 2013). Certain prophage genes can also provide phage immunity via superinfection exclusion (Mahony et al., 2008; Labrie et al., 2010).

The genomes of many C. difficile prophages encode genes that are suspected to influence their host. For example, the large phiCD211-like phages encode putative multidrug resistance genes, spore proteases, and multiple regulators that could interfere with host regulation (Garneau et al., 2018). A CRISPR array with a cas3 gene was also identified, suggesting that phiCD211-like phages possibly participate in CRISPR interference. The presence of CRISPR arrays has been reported in other C. difficile phages, including the two prophages from strain 630 (Hargreaves et al., 2014a; Boudry et al., 2015) as well as the phi027 prophage present in the epidemic strain R20291 and most R027 isolates (Sekulovic and Fortier, 2015). Transcriptomic analyses by RNAseq showed that these CRISPR arrays are transcribed and thus, possibly contribute to C. difficile resistance to invading DNA (Boudry et al., 2015; Sekulovic and Fortier, 2015).

Phage phiCDHM1 and other predicted C. difficile prophages encode homologs of an Agr-like quorum sensing (QS) system (Hargreaves et al., 2014b). QS is used to coordinate specific phenotypes at the whole population level in function of cell density. QS has been implicated in virulence of several pathogens, by coordinating toxin secretion, biofilm production, motility, and sporulation (Novick and Geisinger, 2008; Antunes et al., 2010; Rutherford and Bassler, 2012). The Agr system of Staphylococcus aureus is the most well-characterized QS system in Gram-positive bacteria, and regulates the expression of hundreds of genes, including exotoxins and surface proteins. It is encoded by an operon of four genes, agrD-agrB-agrC-agrA (Novick and Geisinger, 2008). At least two types of QS systems have been described in C. difficile; one is similar to the S. aureus Agr, while the other is related to the luxS/AI-2 from Vibrio harveyi (Stabler et al., 2009). Both systems control the expression of C. difficile toxins in function of cell density (Lee and Song, 2005; Martin et al., 2013; Darkoh et al., 2015). During lysogeny, the agrB and agrC genes are expressed from the phiCDHM1 prophage, suggesting that these components of the QS system are active. However, in the absence of an agrA homolog in phiCDHM1, it is impossible to conclude if the system is functional or not and whether it participates in some way to QS. We can speculate that AgrB and AgrC contribute to autoinducer secretion and signal detection, but no response would be elicited due to the absence of an associated response regulator. Alternatively, these phage-encoded genes could partly complement another Agr system from the host (Hargreaves et al., 2014b). QS likely affects multiple phenotypes in C. difficile (Martin et al., 2013) so it will be interesting to establish whether phage-encoded QS genes influence virulence-associated phenotypes such as toxin production, sporulation, or biofilm formation. QS signals detected by the phage-encoded AgrC could also lead to prophage induction, as observed with soil bacteria (Ghosh et al., 2009). This could be a means for the prophage to “determine” the best moment to initiate a replication cycle that will ensure its successful propagation into the bacterial population. It is therefore reasonable to hypothesize that during infection of the gut, high cell densities would trigger prophage induction, hence promoting phage dissemination and possibly horizontal gene transfer. Further research on the impact of QS on prophage stability would be necessary.

Conclusion and Perspectives

The contribution of phages to the evolution and virulence of C. difficile remains to be clarified (Fortier and Sekulovic, 2013). So far, prophages seem to impact C. difficile’s lifestyle and biology in subtle ways, depending on the genetic background of the host. Studying phage–host interactions requires extensive knowledge of the biology of the phage and the host. Unfortunately, many phage genes have no homologs in databases or have no assigned function. Therefore, one way to investigate the impact of a prophage on its host is to introduce a given temperate phage into a susceptible bacterial host to create a new lysogen and to study various phenotypes in comparison with the parental strain lacking that prophage. However, bacterial genomes often carry multiple prophages and phenotypes can sometimes result from the cumulative effects of more than one prophages, like reported for the DNAses secreted by Streptococcus pyogenes SF370 (Euler et al., 2016). In addition, natural lysogens have been carrying prophages for extensive periods of time and as such, the prophages’ regulatory circuits are often seamlessly integrated into the host network (Ehrbar and Hardt, 2005). Therefore, a better alternative is to remove parts or whole prophages from their natural lysogen to study their impact. Such “prophage-cured” strains can then be compared with the lysogenic parental strain. Curing lysogens from their prophages can be quite challenging depending on the host and the availability of molecular tools. Reports of successful curing using extensive screening for spontaneous prophage-cured mutants, or using allelic exchange with counter selection methods have been published in Gram-negative (e.g., E. coli) and Gram-positive bacteria (e.g., S. pyogenes, S. aureus). These studies have shed light on the role of individual prophages as well as their combined contribution to virulence of their host (Bae et al., 2006; Wang et al., 2010; Euler et al., 2016). Of note, curing of one of the two prophages from C. difficile strain 630 has been recently reported, and involved the use of the CRISPR technology (Hong et al., 2018). This first example of prophage curing in C. difficile paves the way for additional studies on the role of prophages in this pathogen, in particular in epidemic strains such as the R20291 that carries the conserved phi027 prophage (Stabler et al., 2009). Better understanding how phages interact with C. difficile at the molecular level will be essential, especially for future phage therapy applications. Hence research focusing on identifying the cell receptor(s) and the phages’ receptor binding protein and how these two influence the phages’ host range will be crucial. In addition, understanding how phages affect C. difficile and whole bacterial populations in complex ecosystems such as the gut microbiota will be determinant as well. For instance, transfer of certain phages from donors to recipients seems to contribute to the success of fecal microbiota transplantation to treat recurrent C. difficile infections (Zuo et al., 2017). Studying the interplay between the virome and the microbiome in health and disease is thus of high relevance. In conclusion, there is a lot more to discover about C. difficile phages and the newly developed molecular tools and the availability of bacterial genome sequences will certainly foster research in this domain.

Author Contributions

L-CF collected the ideas, concepts, and interpretations, as well as wrote the manuscript.

Funding

This work was supported by a discovery grant from the Natural Sciences and Engineering Research Council of Canada (NSERC; RGPIN-2015-06334). L-CF is a member of the Centre de Recherche du CHUS.

Conflict of Interest Statement

The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

Ackermann, H.-W., and Prangishvili, D. (2012). Prokaryote viruses studied by electron microscopy. Arch. Virol. 157, 1843–1849. doi: 10.1007/s00705-012-1383-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Ainsworth, S., Zomer, A., Mahony, J., and van Sinderen, D. (2013). Lytic infection of Lactococcus lactis by bacteriophages Tuc2009 and c2 triggers alternative transcriptional host responses. Appl. Environ. Microbiol. 79, 4786–4798. doi: 10.1128/AEM.01197-13

PubMed Abstract | CrossRef Full Text | Google Scholar

Amy, J., Bulach, D., Knight, D., Riley, T., Johanesen, P., and Lyras, D. (2018). Identification of large cryptic plasmids in Clostridioides (Clostridium) difficile. Plasmid 9, 25–38. doi: 10.1016/j.plasmid.2018.04.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Antunes, L. C. M., Ferreira, R. B. R., Buckner, M. M. C., and Finlay, B. B. (2010). Quorum sensing in bacterial virulence. Microbiology 156, 2271–2282. doi: 10.1099/mic.0.038794-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Arndt, D., Grant, J. R., Marcu, A., Sajed, T., Pon, A., Liang, Y., et al. (2016). PHASTER: a better, faster version of the PHAST phage search tool. Nucleic Acids Res. 44, W16–W21. doi: 10.1093/nar/gkw387

PubMed Abstract | CrossRef Full Text | Google Scholar

Arndt, D., Marcu, A., Liang, Y., and Wishart, D. S. (2017). PHAST, PHASTER and PHASTEST: tools for finding prophage in bacterial genomes. Brief. Bioinformatics 4:354. doi: 10.1093/bib/bbx121

PubMed Abstract | CrossRef Full Text | Google Scholar

Babu, K. S., Spence, W. S., Monteville, M. R., and Geller, B. L. (1995). Characterization of a cloned gene (pip) from Lactococcus lactis required for phage infection. Dev. Biol. Stand. 85, 569–575.

PubMed Abstract | Google Scholar

Bae, T., Baba, T., Hiramatsu, K., and Schneewind, O. (2006). Prophages of Staphylococcus aureus Newman and their contribution to virulence. Mol. Microbiol. 62, 1035–1047. doi: 10.1111/j.1365-2958.2006.05441.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Bauer, M. P., Notermans, D. W., van Benthem, B. H. B., Brazier, J. S., Wilcox, M. H., Rupnik, M., et al. (2011). Clostridium difficile infection in Europe: a hospital-based survey. Lancet 377, 63–73. doi: 10.1016/S0140-6736(10)61266-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Bebeacua, C., Tremblay, D., Farenc, C., Chapot-Chartier, M.-P., Sadovskaya, I., van Heel, M., et al. (2013). Structure, adsorption to host, and infection mechanism of virulent lactococcal phage p2. J. Virol. 87, 12302–12312. doi: 10.1128/JVI.02033-13

PubMed Abstract | CrossRef Full Text | Google Scholar

Boudry, P., Semenova, E., Sekulovic, O., Fortier, L.-C., Semenov, E., Monot, M., et al. (2015). Function of the CRISPR-cas system of the human pathogen: Clostridium difficile. mBio 6, e1112–e1115. doi: 10.1128/mBio.01112-15

PubMed Abstract | CrossRef Full Text | Google Scholar

Brüssow, H., Canchaya, C., and Hardt, W.-D. (2004). Phages and the evolution of bacterial pathogens: from genomic rearrangements to lysogenic conversion. Microbiol. Mol. Biol. Rev. 68, 560–602. doi: 10.1128/MMBR.68.3.560-602.2004

PubMed Abstract | CrossRef Full Text | Google Scholar

Brüssow, H., and Hendrix, R. W. (2002). Phage genomics: small is beautiful. Cell 108, 13–16.

Google Scholar

Carman, R. J., Stevens, A. L., Lyerly, M. W., Hiltonsmith, M. F., Stiles, B. G., and Wilkins, T. D. (2011). Clostridium difficile binary toxin (CDT) and diarrhea. Anaerobe 17, 161–165. doi: 10.1016/j.anaerobe.2011.02.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Darkoh, C., DuPont, H. L., Norris, S. J., and Kaplan, H. B. (2015). Toxin synthesis by Clostridium difficile is regulated through quorum signaling. mBio 6:e02569. doi: 10.1128/mBio.02569-14

PubMed Abstract | CrossRef Full Text | Google Scholar

Dei, R. (1989). Observations on phage-typing of Clostridium difficile: preliminary evaluation of a phage panel. Eur. J. Epidemiol. 5, 351–354.

PubMed Abstract | Google Scholar

Ehrbar, K., and Hardt, W.-D. (2005). Bacteriophage-encoded type III effectors in Salmonella enterica subspecies 1 serovar Typhimurium. Infect. Genet. Evol. 5, 1–9. doi: 10.1016/j.meegid.2004.07.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Euler, C. W., Juncosa, B., Ryan, P. A., Deutsch, D. R., McShan, W. M., and Fischetti, V. A. (2016). Targeted curing of all lysogenic bacteriophage from Streptococcus pyogenes using a novel counter-selection technique. PLoS One 11:e0146408. doi: 10.1371/journal.pone.0146408

PubMed Abstract | CrossRef Full Text | Google Scholar

Fagan, R. P., and Fairweather, N. F. (2014). Biogenesis and functions of bacterial S-layers. Nat. Rev. Microbiol. 12, 211–222. doi: 10.1038/nrmicro3213

PubMed Abstract | CrossRef Full Text | Google Scholar

Fimlaid, K. A., and Shen, A. (2015). Diverse mechanisms regulate sporulation sigma factor activity in the Firmicutes. Curr. Opin. Microbiol. 24C, 88–95.

Google Scholar

Fortier, L.-C., and Sekulovic, O. (2013). Importance of prophages to evolution and virulence of bacterial pathogens. Virulence 4, 354–365. doi: 10.4161/viru.24498

PubMed Abstract | CrossRef Full Text | Google Scholar

Garneau, J. R., Sekulovic, O., Dupuy, B., Soutourina, O., Monot, M., and Fortier, L.-C. (2018). High Prevalence and Genetic Diversity of Large phiCD211 (phiCDIF1296T)-Like Prophages in Clostridioides difficile. Appl. Environ. Microbiol. 84:e02164-17. doi: 10.1128/AEM.02164-17

PubMed Abstract | CrossRef Full Text | Google Scholar

Gebhart, D., Lok, S., Clare, S., Tomas, M., Stares, M., Scholl, D., et al. (2015). A modified R-type bacteriocin specifically targeting Clostridium difficile prevents colonization of mice without affecting gut microbiota diversity. mBio 6:e02368-14. doi: 10.1128/mBio.02368-14

PubMed Abstract | CrossRef Full Text | Google Scholar

Gebhart, D., Williams, S. R., Bishop-Lilly, K. A., Govoni, G. R., Willner, K. M., Butani, A., et al. (2012). Novel high-molecular-weight, R-type bacteriocins of Clostridium difficile. J. Bacteriol. 194, 6240–6247. doi: 10.1128/JB.01272-12

PubMed Abstract | CrossRef Full Text | Google Scholar

Gerding, D. N., Johnson, S., Rupnik, M., and Aktories, K. (2014). Clostridium difficile binary toxin CDT: mechanism, epidemiology, and potential clinical importance. Gut Microbes 5, 15–27. doi: 10.4161/gmic.26854

PubMed Abstract | CrossRef Full Text | Google Scholar

Ghosh, D., Roy, K., Williamson, K. E., Srinivasiah, S., Wommack, K. E., and Radosevich, M. (2009). Acyl-homoserine lactones can induce virus production in lysogenic bacteria: an alternative paradigm for prophage induction. Appl. Environ. Microbiol. 75, 7142–7152. doi: 10.1128/AEM.00950-09

PubMed Abstract | CrossRef Full Text | Google Scholar

Goh, S., Chang, B. J., and Riley, T. V. (2005a). Effect of phage infection on toxin production by Clostridium difficile. J. Med. Microbiol. 54, 129–135. doi: 10.1099/jmm.0.45821-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Goh, S., Riley, T. V., and Chang, B. J. (2005b). Isolation and characterization of temperate bacteriophages of Clostridium difficile. Appl. Environ. Microbiol. 71, 1079–1083. doi: 10.1128/AEM.71.2.1079-1083.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Goh, S., Hussain, H., Chang, B. J., Emmett, W., Riley, T. V., and Mullany, P. (2013). Phage C2 Mediates Transduction of Tn6215, Encoding Erythromycin Resistance, between Clostridium difficile Strains. mBio 4:e00840-13. doi: 10.1128/mbio.00840-13

PubMed Abstract | CrossRef Full Text | Google Scholar

Goh, S., Ong, P. F., Song, K. P., Riley, T. V., and Chang, B. J. (2007). The complete genome sequence of Clostridium difficile phage phiC2 and comparisons to phiCD119 and inducible prophages of CD630. Microbiology 153, 676–685. doi: 10.1099/mic.0.2006/002436-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Govind, R., and Dupuy, B. (2012). Secretion of Clostridium difficile toxins A and B requires the holin-like protein TcdE. PLoS Pathog. 8:e1002727. doi: 10.1371/journal.ppat.1002727

PubMed Abstract | CrossRef Full Text | Google Scholar

Govind, R., Fitzwater, L., and Nichols, R. (2015). Observations on the Role of TcdE Isoforms in Clostridium difficile Toxin Secretion. J. Bacteriol. 197, 2600–2609. doi: 10.1128/JB.00224-15

PubMed Abstract | CrossRef Full Text | Google Scholar

Govind, R., Fralick, J. A., and Rolfe, R. D. (2006). Genomic organization and molecular characterization of Clostridium difficile bacteriophage PhiCD119. J. Bacteriol. 188, 2568–2577. doi: 10.1128/JB.188.7.2568-2577.2006

PubMed Abstract | CrossRef Full Text | Google Scholar

Govind, R., Vediyappan, G., Rolfe, R. D., Dupuy, B., and Fralick, J. A. (2009). Bacteriophage-mediated toxin gene regulation in Clostridium difficile. J. Virol. 83, 12037–12045. doi: 10.1128/JVI.01256-09

PubMed Abstract | CrossRef Full Text | Google Scholar

Haraldsen, J. D., and Sonenshein, A. L. (2003). Efficient sporulation in Clostridium difficile requires disruption of the sigmaK gene. Mol. Microbiol. 48, 811–821.

PubMed Abstract | Google Scholar

Hargreaves, K. R., and Clokie, M. R. J. (2014). Clostridium difficile phages: still difficult? Front. Microbiol. 5:184. doi: 10.3389/fmicb.2014.00184

CrossRef Full Text | Google Scholar

Hargreaves, K. R., and Clokie, M. R. J. (2015). A Taxonomic Review of Clostridium difficile Phages and Proposal of a Novel Genus, “Phimmp04likevirus”. Viruses 7, 2534–2541. doi: 10.3390/v7052534

PubMed Abstract | CrossRef Full Text | Google Scholar

Hargreaves, K. R., Flores, C. O., Lawley, T. D., and Clokie, M. R. J. (2014a). Abundant and diverse clustered regularly interspaced short palindromic repeat spacers in Clostridium difficile strains and prophages target multiple phage types within this pathogen. mBio 5:e01045-13. doi: 10.1128/mBio.01045-13

PubMed Abstract | CrossRef Full Text | Google Scholar

Hargreaves, K. R., Kropinski, A. M., and Clokie, M. R. J. (2014b). What does the talking?: quorum sensing signaling genes discovered in a bacteriophage genome. PLoS One 9:e85131. doi: 10.1371/journal.pone.0085131

PubMed Abstract | CrossRef Full Text | Google Scholar

Hong, W., Zhang, J., Cui, G., Wang, L., and Wang, Y. (2018). Multiplexed CRISPR-Cpf1-mediated genome editing in Clostridium difficile toward the understanding of pathogenesis of C. difficile Infection. ACS Synth. Biol. 7, 1588–1600. doi: 10.1021/acssynbio.8b00087

PubMed Abstract | CrossRef Full Text | Google Scholar

Horgan, M., O’Sullivan, O., Coffey, A., Fitzgerald, G. F., van Sinderen, D., McAuliffe, O., et al. (2010). Genome analysis of the Clostridium difficile phage PhiCD6356, a temperate phage of the Siphoviridae family. Gene 462, 34–43. doi: 10.1016/j.gene.2010.04.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Kimmitt, P. T., Harwood, C. R., and Barer, M. R. (1999). Induction of type 2 Shiga toxin synthesis in Escherichia coli O157 by 4-quinolones. Lancet 353, 1588–1589.

PubMed Abstract | Google Scholar

Kirk, J. A., Gebhart, D., Buckley, A. M., Lok, S., Scholl, D., Douce, G. R., et al. (2017). New class of precision antimicrobials redefines role of Clostridium difficile S-layer in virulence and viability. Science Transl. Med. 9:eaah6813. doi: 10.1126/scitranslmed.aah6813

PubMed Abstract | CrossRef Full Text | Google Scholar

Labrie, S. J., Samson, J. E., and Moineau, S. (2010). Bacteriophage resistance mechanisms. Nat. Rev. Microbiol. 8, 317–327. doi: 10.1038/nrmicro2315

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, A. S. Y., and Song, K. P. (2005). LuxS/autoinducer-2 quorum sensing molecule regulates transcriptional virulence gene expression in Clostridium difficile. Biochem. Biophys. Res. Commun. 335, 659–666. doi: 10.1016/j.bbrc.2005.07.131

PubMed Abstract | CrossRef Full Text | Google Scholar

Lemire, S., Figueroa-Bossi, N., and Bossi, L. (2011). Bacteriophage crosstalk: coordination of prophage induction by trans-acting antirepressors. PLoS Genet. 7:e1002149. doi: 10.1371/journal.pgen.1002149

PubMed Abstract | CrossRef Full Text | Google Scholar

Mahony, D. E., Bell, P. D., and Easterbrook, K. B. (1985). Two bacteriophages of Clostridium difficile. J. Clin. Microbiol. 21, 251–254.

PubMed Abstract | Google Scholar

Mahony, D. E., Clow, J., Atkinson, L., Vakharia, N., and Schlech, W. F. (1991). Development and application of a multiple typing system for Clostridium difficile. Appl. Environ. Microbiol. 57, 1873–1879.

PubMed Abstract | Google Scholar

Mahony, J., McGrath, S., Fitzgerald, G. F., and van Sinderen, D. (2008). Identification and characterization of lactococcal-prophage-carried superinfection exclusion genes. Appl. Environ. Microbiol. 74, 6206–6215. doi: 10.1128/AEM.01053-08

PubMed Abstract | CrossRef Full Text | Google Scholar

Manrique, P., Bolduc, B., Walk, S. T., van der Oost, J., de Vos, W. M., and Young, M. J. (2016). Healthy human gut phageome. Proc. Natl. Acad. Sci. U.S.A. 113, 10400–10405. doi: 10.1073/pnas.1601060113

PubMed Abstract | CrossRef Full Text | Google Scholar

Manrique, P., Dills, M., and Young, M. J. (2017). The human gut phage community and its implications for health and disease. Viruses 9, 141. doi: 10.3390/v9060141

PubMed Abstract | CrossRef Full Text | Google Scholar

Martin, M. J., Clare, S., Goulding, D., Faulds-Pain, A., Barquist, L., Browne, H. P., et al. (2013). The agr locus regulates virulence and colonization genes in Clostridium difficile 027. J. Bacteriol. 195, 3672–3681. doi: 10.1128/JB.00473-13

PubMed Abstract | CrossRef Full Text | Google Scholar

Mayer, M. J., Narbad, A., and Gasson, M. J. (2008). Molecular characterization of a Clostridium difficile bacteriophage and its cloned biologically active endolysin. J. Bacteriol. 190, 6734–6740. doi: 10.1128/JB.00686-08

PubMed Abstract | CrossRef Full Text | Google Scholar

Meader, E., Mayer, M. J., Gasson, M. J., Steverding, D., Carding, S. R., and Narbad, A. (2010). Bacteriophage treatment significantly reduces viable Clostridium difficile and prevents toxin production in an in vitro model system. Anaerobe 16, 549–554. doi: 10.1016/j.anaerobe.2010.08.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Meader, E., Mayer, M. J., Steverding, D., Carding, S. R., and Narbad, A. (2013). Evaluation of bacteriophage therapy to control Clostridium difficile and toxin production in an in vitro human colon model system. Anaerobe 22, 25–30. doi: 10.1016/j.anaerobe.2013.05.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Meessen-Pinard, M., Sekulovic, O., and Fortier, L.-C. C. (2012). Evidence of in vivo prophage induction during Clostridium difficile infection. Appl. Environ. Microbiol. 78, 7662–7670. doi: 10.1128/AEM.02275-12

PubMed Abstract | CrossRef Full Text | Google Scholar

Modi, S. R., Lee, H. H., Spina, C. S., and Collins, J. J. (2013). Antibiotic treatment expands the resistance reservoir and ecological network of the phage metagenome. Nature 499, 219–222. doi: 10.1038/nature12212

PubMed Abstract | CrossRef Full Text | Google Scholar

Monot, M., Eckert, C., Lemire, A., Hamiot, A., Dubois, T., Tessier, C., et al. (2015). Clostridium difficile: new insights into the evolution of the pathogenicity locus. Sci. Rep. 5:15023. doi: 10.1038/srep15023

PubMed Abstract | CrossRef Full Text | Google Scholar

Mottawea, W., Duceppe, M.-O., Dupras, A. A., Usongo, V., Jeukens, J., Freschi, L., et al. (2018). Salmonella enterica prophage sequence profiles reflect genome diversity and can be used for high discrimination subtyping. Front. Microbiol. 9:836. doi: 10.3389/fmicb.2018.00836

PubMed Abstract | CrossRef Full Text | Google Scholar

Nale, J., Redgwell, T., Millard, A., and Clokie, M. (2018). Efficacy of an optimised bacteriophage cocktail to clear Clostridium difficile in a batch fermentation model. Antibiotics 7:13. doi: 10.3390/antibiotics7010013

PubMed Abstract | CrossRef Full Text | Google Scholar

Nale, J. Y., Chutia, M., Carr, P., Hickenbotham, P. T., and Clokie, M. R. J. (2016a). Biofilm and wax moth (Galleria mellonella) models reveal new insights into the therapeutic potential of Clostridium difficile bacteriophages. Front. Microbiol. 7:1383. doi: 10.3389/fmicb.2016.01383

PubMed Abstract | CrossRef Full Text | Google Scholar

Nale, J. Y., Spencer, J., Hargreaves, K. R., Buckley, A. M., Trzepiński, P., Douce, G. R., et al. (2016b). Bacteriophage combinations significantly reduce Clostridium difficile growth in vitro and proliferation in vivo. Antimicrob. Agents Chemother. 60, 968–981. doi: 10.1128/AAC.01774-15

PubMed Abstract | CrossRef Full Text | Google Scholar

Norman, J. M., Handley, S. A., Baldridge, M. T., Droit, L., Liu, C. Y., Keller, B. C., et al. (2015). Disease-specific alterations in the enteric virome in inflammatory bowel disease. Cell 160, 447–460. doi: 10.1016/j.cell.2015.01.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Novick, R. P., and Geisinger, E. (2008). Quorum sensing in staphylococci. Annu. Rev. Genet. 42, 541–564. doi: 10.1146/annurev.genet.42.110807.091640

CrossRef Full Text | Google Scholar

Oppenheim, A. B., Kobiler, O., Stavans, J., Court, D. L., and Adhya, S. (2005). Switches in bacteriophage lambda development. Annu. Rev. Genet. 39, 409–429. doi: 10.1146/annurev.genet.39.073003.113656

CrossRef Full Text | Google Scholar

Ramírez-Vargas, G., Goh, S., and Rodríguez, C. (2018). The novel phages phiCD5763 and phiCD2955 represent two groups of big plasmidial siphoviridae phages of Clostridium difficile. Front. Microbiol. 9:26. doi: 10.3389/fmicb.2018.00026

PubMed Abstract | CrossRef Full Text | Google Scholar

Rashid, S. J., Barylski, J., Hargreaves, K. R., Millard, A. A., Vinner, G. K., and Clokie, M. R. J. (2016). Two novel myoviruses from the north of iraq reveal insights into Clostridium difficile phage diversity and biology. Viruses 8:310. doi: 10.3390/v8110310

PubMed Abstract | CrossRef Full Text | Google Scholar

Reynolds, C. B., Emerson, J. E., de la Riva, L., Fagan, R. P., and Fairweather, N. F. (2011). The Clostridium difficile cell wall protein CwpV is antigenically variable between strains, but exhibits conserved aggregation-promoting function. PLoS Pathog. 7:e1002024. doi: 10.1371/journal.ppat.1002024

PubMed Abstract | CrossRef Full Text | Google Scholar

Riedel, T., Wittmann, J., Bunk, B., Schober, I., Spröer, C., Gronow, S., et al. (2017). A Clostridioides difficile bacteriophage genome encodes functional binary toxin-associated genes. J. Biotechnol. 250, 23–28. doi: 10.1016/j.jbiotec.2017.02.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Rokney, A., Kobiler, O., Amir, A., Court, D. L., Stavans, J., Adhya, S., et al. (2008). Host responses influence on the induction of lambda prophage. Mol. Microbiol. 68, 29–36. doi: 10.1111/j.1365-2958.2008.06119.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Rupnik, M., Wilcox, M. H., and Gerding, D. N. (2009). Clostridium difficile infection: new developments in epidemiology and pathogenesis. Nat. Rev. Microbiol. 7, 526–536. doi: 10.1038/nrmicro2164

PubMed Abstract | CrossRef Full Text | Google Scholar

Rutherford, S. T., and Bassler, B. L. (2012). Bacterial quorum sensing: its role in virulence and possibilities for its control. Cold Spring Harb. Perspect. Med. 2:a012427. doi: 10.1101/cshperspect.a012427

PubMed Abstract | CrossRef Full Text | Google Scholar

São-José, C., Lhuillier, S., Lurz, R., Melki, R., Lepault, J., Santos, M. A., et al. (2006). The ectodomain of the viral receptor YueB forms a fiber that triggers ejection of bacteriophage SPP1 DNA. J. Biol. Chem. 281, 11464–11470. doi: 10.1074/jbc.M513625200

PubMed Abstract | CrossRef Full Text | Google Scholar

Saujet, L., Pereira, F. C., Henriques, A. O., and Martin-Verstraete, I. (2014). The regulatory network controlling spore formation in Clostridium difficile. FEMS Microbiol. Lett. 358, 1–10. doi: 10.1111/1574-6968.12540

PubMed Abstract | CrossRef Full Text | Google Scholar

Schwan, C., Kruppke, A. S., Nölke, T., Schumacher, L., Koch-Nolte, F., Kudryashev, M., et al. (2014). Clostridium difficile toxin CDT hijacks microtubule organization and reroutes vesicle traffic to increase pathogen adherence. Proc. Nat. Acad. Sci. U.S.A. 111, 2313–2318. doi: 10.1073/pnas.1311589111

PubMed Abstract | CrossRef Full Text | Google Scholar

Schwan, C., Stecher, B., Tzivelekidis, T., van Ham, M., Rohde, M., Hardt, W.-D., et al. (2009). Clostridium difficile toxin CDT induces formation of microtubule-based protrusions and increases adherence of bacteria. PLoS Pathog. 5:e1000626. doi: 10.1371/journal.ppat.1000626

PubMed Abstract | CrossRef Full Text | Google Scholar

Sekulovic, O., and Fortier, L.-C. C. (2015). Global transcriptional response of Clostridium difficile carrying the phiCD38-2 prophage. Appl. Environ. Microbiol. 81, 1364–1374. doi: 10.1128/AEM.03656-14

PubMed Abstract | CrossRef Full Text | Google Scholar

Sekulovic, O., Garneau, J. R., Néron, A., and Fortier, L.-C. C. (2014). Characterization of temperate phages infecting Clostridium difficile isolates of human and animal origins. Appl. Environ. Microbiol. 80, 2555–2563. doi: 10.1128/AEM.00237-14

PubMed Abstract | CrossRef Full Text | Google Scholar

Sekulovic, O., Meessen-Pinard, M., and Fortier, L.-C. (2011). Prophage-stimulated toxin production in Clostridium difficile NAP1/027 lysogens. J. Bacteriol. 193, 2726–2734. doi: 10.1128/JB.00787-10

PubMed Abstract | CrossRef Full Text | Google Scholar

Sekulovic, O., Ospina-Bedoya, M., Fivian-Hughes, A. S., Fairweather, N. F., and Fortier, L.-C. (2015). The Clostridium difficile cell wall protein CwpV confers phase-variable phage resistance. Mol. Microbiol. 98, 329–342. doi: 10.1111/mmi.13121

PubMed Abstract | CrossRef Full Text | Google Scholar

Sell, T. L., Schaberg, D. R., and Fekety, F. R. (1983). Bacteriophage and bacteriocin typing scheme for Clostridium difficile. J. Clin. Microbiol. 17, 1148–1152.

PubMed Abstract | Google Scholar

Shan, J., Patel, K. V., Hickenbotham, P. T., Nale, J. Y., Hargreaves, K. R., and Clokie, M. R. J. (2012). Prophage carriage and diversity within clinically relevant strains of Clostridium difficile. Appl. Environ. Microbiol. 78, 6027–6034. doi: 10.1128/AEM.01311-12

PubMed Abstract | CrossRef Full Text | Google Scholar

Singh, S., Rockenbach, K., Dedrick, R. M., VanDemark, A. P., and Hatfull, G. F. (2014). Cross-talk between Diverse Serine Integrases. J. Mol. Biol. 426, 318–331. doi: 10.1016/j.jmb.2013.10.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Stabler, R. A., He, M., Dawson, L., Martin, M., Valiente, E., Corton, C., et al. (2009). Comparative genome and phenotypic analysis of Clostridium difficile 027 strains provides insight into the evolution of a hypervirulent bacterium. Genome Biol. 10:R102. doi: 10.1186/gb-2009-10-9-r102

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, X., Kim, Y., Ma, Q., Hong, S. H., Pokusaeva, K., Sturino, J. M., et al. (2010). Cryptic prophages help bacteria cope with adverse environments. Nat. Commun. 1:147. doi: 10.1038/ncomms1146

PubMed Abstract | CrossRef Full Text | Google Scholar

Wittmann, J., Riedel, T., Bunk, B., Spröer, C., Gronow, S., and Overmann, J. (2015). Complete Genome Sequence of the Novel Temperate Clostridium difficile Phage phiCDIF1296T. Genome Announc. 3:e00839-15. doi: 10.1128/genomeA.00839-15

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, X., McDaniel, A. D., Wolf, L. E., Keusch, G. T., Waldor, M. K., and Acheson, D. W. (2000). Quinolone antibiotics induce Shiga toxin-encoding bacteriophages, toxin production, and death in mice. J. Infect. Dis. 181, 664–670. doi: 10.1086/315239

PubMed Abstract | CrossRef Full Text | Google Scholar

Zuo, T., Wong, S. H., Lam, K., Lui, R., Cheung, K., Tang, W., et al. (2017). Bacteriophage transfer during faecal microbiota transplantation in Clostridium difficile infection is associated with treatment outcome. Gut 67, 634–643. doi: 10.1136/gutjnl-2017-313952

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: Clostridium difficile, Clostridioides difficile, bacteriophages, prophages, toxins, virulence, lysogenic conversion

Citation: Fortier L-C (2018) Bacteriophages Contribute to Shaping Clostridioides (Clostridium) difficile Species. Front. Microbiol. 9:2033. doi: 10.3389/fmicb.2018.02033

Received: 30 June 2018; Accepted: 13 August 2018;
Published: 31 August 2018.

Edited by:

Meina Neumann-Schaal, Deutsche Sammlung von Mikroorganismen und Zellkulturen (DSMZ), Germany

Reviewed by:

Imrich Barak, Institute of Molecular Biology (SAS), Slovakia
Shan Goh, University of Hertfordshire, United Kingdom

Copyright © 2018 Fortier. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Louis-Charles Fortier, Louis-Charles.Fortier@USherbrooke.ca

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.