Skip to main content
FEMS Microbiology Reviews logoLink to FEMS Microbiology Reviews
. 2022 Jul 26;46(6):fuac036. doi: 10.1093/femsre/fuac036

Barriers to genetic manipulation of Enterococci: Current Approaches and Future Directions

Alexandra L Krause 1, Timothy P Stinear 2, Ian R Monk 3,
PMCID: PMC9779914  PMID: 35883217

Abstract

Enterococcus faecalis and Enterococcus faecium are Gram-positive commensal gut bacteria that can also cause fatal infections. To study clinically relevant multi-drug resistant E. faecalis and E. faecium strains, methods are needed to overcome physical (thick cell wall) and enzymatic barriers that limit the transfer of foreign DNA and thus prevent facile genetic manipulation. Enzymatic barriers to DNA uptake identified in E. faecalis and E. faecium include type I, II and IV restriction modification systems and CRISPR-Cas. This review examines E. faecalis and E. faecium DNA defence systems and the methods with potential to overcome these barriers. DNA defence system bypass will allow the application of innovative genetic techniques to expedite molecular-level understanding of these important, but somewhat neglected, pathogens.

Keywords: Enterococcus faecium, Enterococcus faecalis, restriction modification, electroporation, molecular biology


The molecular biology of Enterococcus faecalis and Enterococcus faecium has lagged behind other Gram-positive pathogens due to physical and enzymatic barriers restricting DNA transfer, with this review summarising the barriers and highlighting how improving transformation will accelerate our understanding of these important pathogens.

Introduction

The Gram-positive genus Enterococcus comprises over 50 species of commensal gastrointestinal (GI) bacteria that make up about 1% of the total microbial load in healthy individuals, but some species of the genus are also a major cause of nosocomial infections (causing 14% of all hospital acquired infections in the US between 2011 and 2014) (Fiore et al.2019a). They can inhabit various niches including different regions of the GI tract and oral cavity of humans and other animals, but they can also be recovered from water, soil, plants and processed food (Franz et al. 2003, Byappanahalli et al. 2012, Hammerum 2012, Lebreton et al. 2014, Komiyama et al. 2016, Dubin et al. 2017). The most notable enterococcal human pathogens are Enterococcus faecalis and Enterococcus faecium. When pathogenic enterococci gain access to niches in the human body beyond their commensal habitat of the GI tract, such as the urinary tract, skin (via wounds/burns), bloodstream and the heart, they have a high potential to cause persistent infections and subsequent serious disease (Fiore et al.2019a). These types of infections are more common in people with underlying risk factors that include immunosuppression, renal/liver disease, diabetes mellitus, or abdominal and oral surgery (Billington et al. 2014, Kajihara et al. 2015).

Enterococcal infections, in particular those caused by E. faecalis and E. faecium, can be very problematic because of the propensity of the bacteria for intrinsic and acquired multiple drug resistance (MDR). E.faecium infections can be very difficult to treat because of high levels of ampicillin and vancomycin resistance, compared to E. faecalis where ampicillin and vancomycin resistance are less frequently reported. Antibiotics such as linezolid, daptomycin, tigecycline, and quinupristin-dalfopristin are commonly used to treat MDR enterococcal infections, however, resistance to these drugs has also been observed (Kristich et al. 2014, Mercuro et al. 2018). Analysis of E. faecium clinical isolates found that 13.4% (106/791) compared with 1.4% (9/651) E. faecalis strains were resistant to linezolid (Malisova et al. 2021). A cohort study collected E. faecium isolates from patients colonised with vancomycin resistant enterococci (VRE) and found that 11.7% (50/426) of the strains examined were also resistant to daptomycin (Kinnear et al. 2020), observations that are in line with a resistance rate (15%–49/324) recently reported from a collection of Australian and New Zealand E. faecium isolates (Li et al. 2021). E. faecalis is intrinsically resistant to quinupristin-dalfopristin conferred by the lsa gene, with resistance in E. faecium also mediated by drug efflux (Hollenbeck and Rice 2012). E. faecium resistance to quinupristin-dalfopristin was 1% (9/911) in a study from China (Wang et al. 2016), but 9.4% (16/170) in Australia (Coombs et al. 2013). Tigecycline resistance is reported to only be 1% for E. faecium and 0.3% for E. faecalis, which might be due to the bacteriostatic mode of action of the drug (Dadashi et al. 2021). In agriculture, the use of antibiotics including tetracycline, streptomycin, avoparcin (a glycopeptide) and phenicols (often though co-location on the chromosome with optrA or cfr conferring linezolid resistance) are a major driver of antibiotic resistance in enterococci (Manson et al. 2004, Kristich et al. 2014, Nuesch-Inderbinen et al. 2021, Timmermans et al. 2021). To combat the MDR problem, newer antibiotics such as dalbavancin, oritavancin, tedizolide, and telavancin are being used but with varying degrees of success and availability (Kristich et al. 2014, Esmail et al. 2019, Krawczyk et al.2021).

When the GI tract of a patient is colonised with VRE, the use of pre- (or post) operative antibiotics can cause VRE to become the dominant species in the gut microbiota and then spread to other body sites (Fiore et al. 2019b, Krawczyk et al.2021). MDR strains of VRE are an increasing burden on the healthcare system. Transient carriage of MDR VRE on the hands of healthcare workers can be a major transmission vector and increase hospital infections if not properly cleaned and sanitised (Grayson et al. 2012, Agudelo Higuita and Huycke 2014, Krawczyk et al.2021) with some E. faecium strains also exhibiting an increased tolerance to alcohol (Pidot et al. 2018).

E. faecium strains can be divided into subspecies made up of clades A and B, typically clade A are MDR E. faecium strains and clade B are commensal, non-hospital, antibiotic susceptible strains, which have been recently renamed E. lactis. (Belloso Daza et al. 2021). A further subspecies divergence has been identified among the clades into A1 and A2. Clade A1 are linked to more infective, endemic hospital strains, while clade A2 are linked to periodic infection in humans from an animal reservoir (Lebreton et al. 2013). With increased genomic resolution afforded through the analysis of a larger E. faecium strain collection from the UK and Ireland, the original A1/A2 separation was not supported. However, A2 clade isolates were shown to be ancestorial to isolates from A1 (Raven et al. 2016). A further in-depth analysis of a representative collection of international E. faecium isolates, suggested that a continuum exists between the A and B clades, with A comprising the dominant hospital lineage (van Hal et al. 2022). Further subgroups, approximately equivalent to the clade structure of A1 and A2, identified by Lebreton et al, could be distinguished (Lebreton et al. 2013). The authors suggest that these can be missed if pre-hospital colonising (predominantly A2 and B) and hospital acquired isolates (A1) are not adequately sampled. E. faecium is highly recombinogenic, with gene flow observed from A2 or B isolates to A1 isolates. Each group contains a unique repertoire of genes allowing for the generation of enhanced genetic diversity through gene introduction or formation of hybrid genomes (van Hal et al. 2022). Additionally, E. faecium contains extensive diversity in their plasmid content, with the complete plasmidome of 62 clade A isolates recently described, showing the presence of 305 plasmids. Interestingly, E. faecium isolates from hospitalised patients contained a significantly increased plasmid DNA content. Optimal plasmid configurations were potentially important in improving strain fitness in the hospital environment (Arredondo-Alonso et al. 2020). In contrast, limited phylogenetic distribution has been observed in E. faecalis, without an obvious lineage or clade structure (Palmer et al. 2012).

The presence of restriction modification (RM) systems leading to host specific DNA methylation signatures are major hurdles that prevent detailed molecular studies of these opportunistic pathogens. Most of the fundamental research on the biology and pathogenesis of E. faecalis and E. faecium has been conducted with select strains able to be transformed efficiently with plasmids isolated from Escherichia coli (Jacob and Hobbs 1974, Paulsen et al. 2003, Nallapareddy et al. 2006b, Bourgogne et al. 2008, Zhang et al. 2012). However, these strains are not representative of most clinical isolates (Nallapareddy et al.2006a, Huo et al. 2015, de Maat et al. 2020), where physical and RM barriers prevent the introduction of foreign DNA. The inability to readily and reliably introduce DNA, to conduct targeted or random mutagenesis in E. faecalis and E. faecium represents a major bottleneck for research efforts to understand their pathobiology. In this review, we will detail the physical and enzymatic barriers present on the outside/inside of the enterococcal cell, respectively, with a specific focus on E. faecium. Progress in breaching these barriers will facilitate the application of new genetic tools and accelerate progress towards an improved understanding of E. faecium at a molecular level.

Gram-positive cell wall structure; a barrier to foreign DNA uptake

The thick enterococcal cell wall (approximately 40 nm) is composed of peptidoglycan (PG), interspersed with wall polymers such as wall teichoic acid (WTA) or other polysaccharides. Lipoteichoic acids (LTA) extend from the top of the peptidoglycan layer and is anchored to the plasma membrane by covalent bonds (Fig. 1) (Shockman and Barrett 1983, Hancock et al. 2014, Brown et al. 2015, Yang et al. 2017, Chang et al. 2018). The cell wall also contains cell wall polysaccharides such as capsular polysaccharides (Cps) which are embedded in the cell wall and extend outwards, as well as enterococcal polysaccharide antigens (EPA) (Hancock et al. 2014).

Figure 1.

Figure 1.

The enterococcal cell wall architecture. The cell wall of enterococci is composed of a thick peptidoglycan cross-linked lattice of repeating sugars (N-acetylmuramic acid (NAM) and N-acetylglucosamine (NAG)) containing wall teichoic acids, lipoteichoic acids, capsular polysaccharides and enterococcal polysaccharide antigen (EPA) polymers surrounding a plasma membrane made up of a phospholipid bilayer. The cell wall can also contain membrane bound (lipoprotein shown) or cell wall bound proteins (sortase A linked PrgB and TraO surface adhesions are shown). The impact of glycine overload on peptidoglycan crosslinking is detailed (Shockman and Barrett 1983, Hancock et al. 2014, Brown et al. 2015). Created with BioRender.com.

The recent in-depth review from Ramos et al. summarises the roles of enterococcal surface polysaccharides in the host-pathogen interaction (Ramos et al. 2021). The main component of the cell wall is PG and this forms a rigid repeating lattice structure of glycans (N-acetylmuramic acid and N-acetylglucosamine). These sugars are linked together by glycosidic bonds catalysed by a transglycosylase. (Shockman and Barrett 1983). Schleifer and Kilpper-Balz (Schleifer and Kilpper-Bälz 1987) showed that of 7 out of 8 enterococcal species examined contain a peptide cross bridge comprising of a single D-Asp, whereas in E. faecalis it is comprised of 2–3 L-Ala residues. In E. faecium cross bridge location on the linking residues can be altered by environmental conditions (Ampicillin or Vancomycin exposure) and genetic background (VSE vs VRE). The substrate for penicillin binding proteins (D,D-transpeptidase) is the pentapeptide stem, which catalyses 4 (D-Ala)-3 (L-Lys) crossbridge (D-Asp or L-Ala2-3 formation). Alternately, in the presence of tetrapeptide substrate which can occur through two independent mechanisms depending on the antibiotic, a L,D-transpeptidase can catalyse 3(L-Lys)-3(L-Lys) cross-bridge formation (Pidgeon et al. 2019). These strong bonds unify the PG in the cell wall making it very hard to penetrate, even with electroporation, without first weakening the cell wall to get DNA in the cell (Aune and Aachmann 2010). Glycine added to growth media is commonly used to weaken the cell wall of Gram-positive bacteria as it can substitute a L-alanine or D-alanine in the stem peptide (Fig. 1) that subsequently reduces the strength of the cross linking (Hammes et al. 1973). Cell wall weakening occurs as the altered glycine-containing PG precursors are often not cross-linked by a transpeptidase, which can lead to changes in cell morphology (Hammes et al. 1973, Bhattacharjee and Sorg 2020). This approach has been successfully applied to the generation of electrocompetent cells in a wide array of Gram-positive bacteria (Holo and Nes 1989, Buckley et al. 1999, Turgeon et al. 2006, van Pijkeren et al. 2010, Chang et al. 2017, Bhattacharjee and Sorg 2020). Alterations to the terminal pentapeptide stem residues from D-Ala-D-Ala to D-Ala-D-Lac in VRE confers vancomycin resistance due to reduced target binding affinity (Hancock et al. 2014), while the impact of this cell wall modification on transformation has not been determined. Overall, the thick rigid lattice structure of the peptidoglycan layer in the cell wall poses a substantial barrier to DNA uptake without any other mechanism of natural DNA competence present in enterococci.

A second principle component of the cell wall are the essential teichoic acids, of which there are two types of anionic glycopolymers; WTA and LTA (Theilacker et al. 2012). In enterococci teichoic acids have roles in host cell interaction, biofilm formation, resistance to antimicrobial peptides and also cell wall cation homeostasis (Theilacker et al. 2012, Ramos et al. 2021). In Gram-positive bacteria the tagO (also called tarO) gene is the first committed step of WTA biosynthesis (Holland et al. 2011, Palmer et al. 2012, Brown et al. 2013, Winstel et al. 2013) with the enterococcal ortholog being epaA (Palmer et al. 2012). In S. aureus, in addition to other roles, WTA acts as temporal and spatial regulator of peptidoglycan metabolism (Atilano et al. 2010). Deletion of tagO has pleiotropic effects (impaired cell division, reduced PG cross-linking, protein localisation and virulence) in all bacteria examined, with the deletion from Staphylococcus epidermidis also enhancing transformation by electroporation, indicating a potential role of WTA composition in limiting horizontal gene transfer (HGT) (Atilano et al. 2010, Holland et al. 2011). WTA serves as a receptor for certain Staphylococcus aureus phages, variation in WTA glycosylation can impact on phage host range (Winstel et al. 2013).

The presence of a capsule has been shown to reduce DNA uptake in a number of bacterial pathogens. In E. faecalis it plays a role in host immune evasion and resistance to opsonophagocytosis (Fournet-Fayard et al. 1995, Jeon et al. 2009, Thurlow et al. 2009, Brown et al. 2013). Additionally, mutations in capsule biosynthesis are likely to have an impact on the integrity of the enterococcal cell wall, which could be beneficial for transformation (Ali et al. 2019). EPA is another conserved enterococcal cell wall antigen which is composed of repeating sugar subunits that can vary between enterococcal strains and has been shown to be involved in immune evasion, intestinal colonisation and virulence (Chatterjee et al. 2019, Smith et al. 2019). When EPA is mutated, it is shown to decrease peptidoglycan crosslinking and impair cell wall integrity (Smith et al. 2019). Currently, the impact of EPA on enterococcal electroporation is unknown, but it plays an essential role in the adsorption and infection by lytic bacteriophages in E. faecium and E. faecalis (Chatterjee et al. 2019, Canfield and Duerkop 2020, Canfield et al. 2021).

A number of publications have systematically analysed the cell wall as a barrier in specific Gram-positive bacteria to develop optimal treatments to achieve maximal DNA transfer by electroporation (Aune and Aachmann 2010). While the cell wall of E. faecalis and E. faecium provides a formidable barrier, strains can also encode a strong second line of defence, as discussed below (Manso et al. 2014, Lee et al. 2019).

DNA defence systems in E. faecium and E. faecalis

The transfer of DNA between bacteria encourages evolution by the introduction and incorporation of new genetic material. However, there is a balance between essential and superfluous genes and the need for bacteria to maintain a streamlined genome to be competitive in complex microbiomes (Oliveira et al. 2014). There are two main bacterial defense mechanisms that prevent HGT, either clustered regularly interspaced short palindromic repeats and the associated proteins (CRISPR-Cas) (Nidhi et al. 2021) or RM systems (Murray 2000).

CRISPR-Cas function as a bacterial ‘immune system’ that evolves through integration of DNA recognised as foreign into a CRISPR array, which then become CRISPR guide RNAs. The guides target foreign DNA to be specifically degraded by nuclease encoded activity. The presence of CRISPR-Cas is generally more common in commensal enterococcal strains which prevent the uptake of foreign DNA such as the acquisition of drug resistance elements (Johnson et al. 2021). Conversely, MDR/clinical enterococcal strains either lack these defence systems or contain non-functional variants such as CRISPR2 (contains spacer array but no cas genes) (Johnson et al. 2021). Therefore, the MDR strains have increased genetic plasticity which can enhance survival compared to strains that lack genetic diversity due to the absence of MDR genes (Johnson et al. 2021). This is exemplified in E. faecium, where in clade B strains (commensal strains) examined, CRISPR-Cas were sporadically present, with few MGE. While clades A1/A2, no CRISPR-Cas genes were identified and the genome were littered with antimicrobial resistance (AMR) determinants (Lebreton et al. 2013). However, this simplistic view does not take into account the presence of other barriers discussed below. A recent comprehensive genomic analysis of all available genome sequences identified a greater prevalence of CRISPR spacers in E. faecalis strains (∼75%) compared to ∼5% in E. faecium strains (Mlaga et al. 2021), but this did not assess whether the putative systems identified were functional.

The second defence mechanism is RM, with only three out of the four known RM systems described in E. faecium and E. faecalis (the type III RM systems have not been detected).

Type I RM systems are the most complex and are made up of three protein subunits, termed host specificity of DNA (Hsd), specificity (S) protein, the modification (HsdM) (methylation) protein and the restriction endonuclease (HsdR) protein (Monk and Foster 2012). HsdM and HsdS form a protein complex (HsdM2 HsdS1) yielding the adenine methylase activity. The complex binds to a specific DNA sequence dictated by the two-target recognition domains (TRD) of the HsdS. These DNA motifs are comprised of an asymmetric bipartite DNA sequence with the two TRD DNA motifs comprising 3–4 base pairs (containing the methylated adenine residue) separated by 4 to 9 non-specific base pairs. In the presence of TRD specific adenine methylation (on hemi-methylated replicating DNA), methylation occurs at the corresponding unmethylated TRD site. The addition of two HsdR subunits to the HsdM2HsdS1 complex at completely unmethylated sites transforms the system into a molecular motor and it is translocated along flanking DNA via ATP hydrolysis, breaking when collision with either a second bound RM complex or DNA secondary structure. Therefore, Type I RM systems do not yield a defined cleavage sequence pattern. The formation of the restriction complex and subsequent DNA cleavage is inhibited by the correct adenine methylation profile.

Type II RM systems are a diverse collection of enzymes with sequence specific DNA cleavage, with examples including type IIP EcoRI or type IIS BsaI. Cleavage can be prevented or activated by the presence or absence of specific methylation on the recognition sequence. These enzymes have been the ‘workhorses’ in molecular biology, with their discovery and application have dramatically accelerated the application of recombinant DNA technologies in biotechnology (Roberts 2005). A comprehensive review of type II RM systems was published in 2014 (Pingoud et al. 2014).

Type IV RM systems are the simplest and function by restriction enzyme cleavage of modified DNA e.g. methylated or glycosylated DNA. They lack methyltransferase activity, with the system operating inversely to the other RM systems (Loenen et al. 2014). Type IV systems have been identified across a broad array of Gram-positive and Gram-negative bacteria, where they present a major barrier to prevent foreign DNA uptake (Gonzalez-Ceron et al. 2009, Xu et al. 2011, Sitaraman 2016, Gifford et al. 2019, Woods et al. 2019, Picton et al. 2021).

PacBio (a DNA sequencing platform that can specifically discriminate methylated DNA bases) sequencing (Flusberg et al. 2010, Clark et al. 2012) has helped identify the contribution of RM in E. faecium and E. faecalis to the development of the distinct genetic clades (Furmanek-Blaszk and Sektas 2015, Huo et al. 2015). Sequencing confirms that clinical isolates like E. faecium strains 6E6 and 2014-VREF-63 have 6-methyl adenine DNA methylation, which is associated with a type I RM system and 5-methyl cytosine, which is associated with a type II RM system (Huo et al. 2019).

E. faecium isolates also harbour a type IV RM system that is overrepresented in B and A2 clades (Lebreton et al. 2013), while clade A1 clade strains are enriched for a specific type I RM system (Huo et al. 2019). It has been suggested that these DNA defence systems could be a driver for the formation of distinct clades found in E. faecium and potentially contributing to the issues with genetic manipulation of these strains (Lebreton et al. 2013). Comparative genomic analysis of 73 E. faecium strains (15 clade B, 21 clade A1, 35 clade A2 and 2 hybrid isolates) identified a conserved type I system present in clade A1, with specific hsdM and hsdR alleles identified in 18/21 isolates (Lebreton et al. 2013). A single HsdS allele was found in 14/21 clade A1 isolates, which may have contributed to A1 subspeciation and selected for phenotypes with improved survival characteristics in healthcare settings. However, unlike in S. aureus, where up to 100% of a clonal complex contain a specific HsdS repertoire (Lee et al. 2019), the single type I system in clade A1 E. faecium might yield a partial barrier to DNA transfer. It will be important to screen more E. faecium genome sequences to assess the distribution of this particular hsdS allele. Interestingly, after deletion of the type I RM system (ΔRM) from two different E. faecium backgrounds (clade A1; strain 1231 502 and clade B; strain 1141 703) there was only up to a half log increase in transformation efficiency between the ∆RM mutant compared to the wild type (Huo et al. 2019). This questions the impact this RM system has on foreign DNA transfer. However, the test plasmid used for the electroporations contained only one HsdS recognition sequence. A plasmid containing multiple recognition sequences would help improve the resolution and clarify the impact of type I RM in limiting DNA transfer. A more recent take on the presence of type I RM related to plasmid content identified the same HsdS allele described by Huo et al (Huo et al. 2019) as being enriched, but also described an additional eight HsdS alleles (Arredondo-Alonso et al. 2020). Of these eight, four were found in A1 and two in non-A1 isolates. Authors suggest that type I RM may play a role in shaping the plasmid content of E. faecium clones through the restriction of DNA transfer (Arredondo-Alonso et al. 2020).

In E. faecalis type I RM systems are also present and can be found as phasevarions (phase-variable regulons); an epigenetic regulatory system (Atack et al. 2020a). Phase variation is the random switching on and off of gene expression at a high frequency, this typically occurs in genes encoding surface proteins such as pili or adhesins of host-adapted pathogens (Phillips et al. 2019a). The presence of a potential RM system in E. faecalis was first postulated in 1978 after plasmid DNA isolated from E. faecalis had a lower transformation efficiency compared to DNA isolated from S. sanguis. (Leblanc et al. 1978). The SfeI a type II RM system subsequently discovered in E. faecalis, is closely related to an ortholog from Lactococcus lactis (Okhapkina et al. 2002). Recently, two more type II RM systems have been discovered in E. faecalis OG1RF and E. faecalis NEB215 (Huo et al. 2015).

Phase variation can occur through simple sequence repeats (SSRs) located in an open reading frame or promoter region, leading to frameshift mutations or an impact on the expression level, respectively. Additionally, SSRs can be affected by length of the SSRs which can cause fluctuation in gene expression levels. Phase variation of type I RM through modulation of methyltransferase hsdM expression can lead to methylation differences across the genome resulting in a global impact on gene expression (Atack et al. 2020a,b). It was determined that around 10% of type I RM systems contain SSRs and can undergo phase variation of gene expression (Atack et al. 2020a). Invertase mediated recombination through TRD of co-localised HsdS alleles can change the specificity of the HsdS allele or inactivate the protein. When there are multiple HsdS proteins these can create complex phasevarions. These complex methylation patterns provide heterogeneity in gene expression which is akin to a bet-hedging strategy which can improve fitness under certain conditions (Reyes Ruiz et al. 2020).

The use of phase-regulated gene expression allows the bacteria to encode multiple genes that are beneficial for adapting to different environments e.g. evading immune response, colonisation of a different host niche or surviving in a harsh environment (Phillips et al. 2019a). Atack et al (Atack et al. 2020a), identified that 24/34 E. faecalis strains analysed contained a type I RM system. The 24 type I RM systems encoded multiple hsdS genes, with a high level of TRD variability. This showed that phase variable type I RM systems are present in E. faecalis, and begs the question if as yet unidentified phasevarions are also present in E. faecium (Phillips et al. 2019a, Huang et al. 2020, Atack et al. 2020a).

In S. aureus, both naturally occurring and constructed restriction deficient strains have been essential tools in unlocking S. aureus genetics (Kreiswirth et al. 1983, Monk and Foster 2012). These restriction deficient strains have the ability to accept large drug resistance plasmids from E. faecalis by conjugation, potentially contributing to the dissemination of vancomycin resistance genetic elements into S. aureus population (Sung and Lindsay 2007). This work highlighted that the restriction barrier can be overcome by using strains lacking fully functioning RM systems, whereby they are methylation proficient but restriction deficient. Further work has been conducted in S. aureus along these lines to streamline genetic manipulation. An E. coli strain, DC10B, was constructed through deletion of the dcm gene to prevent cytosine methylation of DNA and subsequent degradation by the conserved S. aureus type IV RM system (Xu et al. 2011). The combination of DC10B with an improved allelic exchange vector (pIMAY—pWV01ts and strong antibiotic selection) has streamlined construction of S. aureus and S. epidermidis mutants (Monk et al. 2012, Monk and Stinear 2021). Further E. coli and plasmid improvements have accelerated the ability to engineer S. aureus strains. Clonal complex specific type I RM methylation profiles were heterologously expressed from the chromosome of DC10B which allowed the complete bypass type I and type IV RM. The addition of blue/white screening into pIMAY (yielding pIMAY-Z) established phenotypic screening of plasmid loss (Monk et al. 2015). Similar plasmid modification techniques could be applied to E. faecium strains containing type I RM, as pIMAY-Z has already been successfully used in E. faecium (Pidot et al. 2018, Monk and Stinear 2021). We have found that the Prokaryotic Antiviral Defence Locator (https://padloc.otago.ac.nz/padloc/) bioinformatic tool as a rapid and user friendly method to assess the presence of RM systems and other barriers, which could impair plasmid transfer in enterococci (Payne et al. 2021).

Horizontal gene transfer in enterococci

Certain bacteria such as Haemophilus influenzae, Neisseria gonorrhoeae, Bacillus subtils, and Streptococcus pneumoniae can actively uptake environmental DNA through the production type IV pili which pull DNA into the cell (Giltner et al. 2012, Mell and Redfield 2014, Muschiol et al. 2015). Under activating conditions cells become naturally competent with cues including cell density, presence of specific nutrients and induction of stress networks (Mell and Redfield 2014, Muschiol et al. 2015, Piepenbrink 2019). Type IV pili for DNA uptake have not been identified in enterococci. Other naturally occurring systems of DNA transformation and HGT are present in enterococci including conjugation and transduction, which are detailed below.

For bacteria which lack these natural competence systems, an ‘un-natural’ approach to DNA transformation is required to transfer recombinant DNA into the cell. One widely used approach involves the application of electric currents called electroporation, to induce transient pores in the bacteria cell wall and allow DNA to enter (Aune and Aachmann 2010). Prior to the development of electroporation, a common technique for HGT was protoplast transformation. PG was enzymatically digested which facilitated DNA uptake, with subsequent cell wall regeneration (Wirth et al. 1986). The natural HGT systems of conjugation and transduction can also be manipulated to allow for the targeted genetic modification.

Conjugation

The transfer of DNA from a donor bacterial cell to a recipient cell which requires close contact and generation of a pilus/pore is called conjugation. This mechanism of transfer is common among plasmids and integrative conjugative elements (ICEs) harbouring AMR genes (Perry and Wright 2013, Kohler et al. 2019). However, in Gram-positive bacteria conjugation has not been extensively used as a tool for genetic manipulation as it relies on the bacteria expressing complex machinery to carry out conjugation (Type IV secretion system) and also has to overcome host enzymatic barriers such as CRISPR and RM (Dunny 2007, Strand et al. 2014). In E. faecalis, pheromone-inducible plasmid transfer is a common mechanism of plasmid transfer, with it less frequently observed in E. faecium (Sterling et al. 2020). This involves the secretion of pheromones by the recipient to the donor cells stimulating the production of aggregation proteins such as PrgB, which initiates close cell-to-cell contact, formation of a mating channel and then plasmid transfer (Clewell et al. 1985, Dunny et al. 1995, Sterling et al. 2020). Enterococci share ssDNA efficiently via conjugation. Conjugative plasmids can encode anti-restriction genes, like ardA. ArdA mimics dsDNA and inhibits type I RM activity allowing ssDNA to dsDNA conversion. (Marcinek et al. 1998, Murray 2000, McMahon et al. 2009, Zavil'gel'skii and Rastorguev 2009).

Pheromone inducible plasmids like tetracycline-resistance plasmid pCF10 contain a specific Prg/Pcf conjugation system. These respond only to CF10 pheromone levels through a quorum sensing mechanism to initiate conjugation (Dunny 2007). Pheromone inducible plasmids are ideal tools for genetic manipulation as they are highly efficient at DNA transfer (Staddon et al. 2006, Kohler et al. 2019). Enterococcal conjugative plasmids can additionally mobilise other plasmids in the cell (even non-conjugative plasmids) and lead to the formation of chimeric plasmids through recombination (Di Sante et al. 2017). This can be an issue when working with a clinical strain which contains multiple plasmids (Shan et al. 2022). Enterococcal conjugative plasmids containing hylEfm have been shown to co-transfer with antibiotic resistance regions such as vanA and may play a role in the dissemination of antibiotic resistance elements in clinical E. faecium strains (Arias et al. 2009).

Another important plasmid family for E. faecalis and E. faecium are Inc18-containing plasmids, e.g. pIP501, which encode the tra operon (traA, traN and traO), that controls the conjugative processes. TraA is a relaxase that negatively regulates tra-operon transcription. TraN has a multi-layered function and works as a repressor of the conjugative system by blocking the tra-operon promoter (Ptra) and also represses the TraNO promoter (PtraNO). TraN is postulated to regulate itself as well as production of TraO, the surface adhesion required for cell-to-cell conjugation (Kohler et al. 2019). Conjugation protocols for enterococci are based on the single stationary phase timepoint protocol described by Simonsen et al(Simonsen et al. 1990). The recipient cell is resistant to an antibiotic marker not produced by the donor and therefore allowing for the direct selection of transconjugants upon plasmid transfer. Clinical E. faecalis and E. faecium strains are normally spoilt for choice due to the presence of multiple antibiotic markers. Kristich et al. constructed a spectinomycin-resistant E. faecalis donor strain for RepA-dependent conjugative plasmid delivery (pWV01 replicon plasmids), called CK111 (Kristich et al. 2007b). The chromosomally integrated pWV01 repA gene is under the control of the lactococcal P23 promoter, which was inserted into the upp locus, making the donor strain resistant to 5-flourouracil (5-FU). E. faecalis CK111 would be ideal for use in plasmid transfer experiments into clinical E. faecium strains.

Bacteriophage transduction

Another mechanism of HGT is transduction, which is the introduction and potential integration of DNA into the host cell by a bacteriophage or virus like particles (VLPs) (Kleiner et al. 2020). Temperate bacteriophage enter the cell and multiply without disrupting the cells, while lytic phages multiply and cause cell lysis. Temperate phages can integrate at specific attB site(s) in the host genome and enter a lysogenic phase of phage dormancy. Integrated phage can excise from the chromosome and replicate upon sensing appropriate signal(s) (Mazaheri Nezhad Fard et al. 2011, Johnson et al. 2021).

Enterococcal bacteriophages are heavily implicated in HGT of antimicrobial and virulence genes in bacterial populations. Clinical MDR E. faecium have twice as many MGE-associated genes as non-clinical strains, which have most likely been acquired via MGE such as bacteriophages or plasmids (Buultjens et al. 2017, Johnson et al. 2021). E. faecalis VRE isolate V583 contains seven prophage-like elements that could contribute to the virulence and fitness of this isolate (Johnson et al. 2021). However, even though MDR E. faecium strains pose a threat to healthcare little is known about the role bacteriophage play in the population (eg. acquisition of novel virulence, immune evasion or antibiotic resistance genes), due to research being hindered by issues with current genetic manipulation techniques (Mazaheri Nezhad Fard et al. 2011, Johnson et al. 2021).

Generalised transduction is the non-specific accidental packaging of DNA into a bacteriophage during replication, which can facilitate the transfer of plasmids or chromosomal loci from one bacteria to another (Chiang et al. 2019). The application of generalized DNA transduction was first documented in Salmonella with phage P22, which is a pac-type phage which recognise pac sites on the bacterial or plasmid DNA to start DNA packing into the phage (Zinder and Lederberg 1952, Chiang et al. 2019). Lysogenic phages of E. faecalis have been identified which are capable of generalised transduction, which potentially could be applied to E. faecium (Yasmin et al. 2010).

Specialised transduction occurs through the accidental transfer of sequence specific DNA in proximity to the chromosomal phage attachment site and subsequent transfer. An example of a specialized transducing phage is E. coli lambda, which can also co-transduce with gal locus, at its integration site (Rédei 2008). In Mycobacterium tuberculous specialised transduction has been repurposed to facilitate a highly efficient one step allelic exchange system (Jacobs 2014, Jain et al. 2014). A similar approach could be modified and used to alter gene expression or construct targeted mutations in E. faecium. The presence of lysogenic bacteriophage could be identified through mitomycin C treatment of E. faecium strains with their ability to conduct generalised or specialised transduction assayed. Additionally, phage integrase genes could be repurposed for the construction of single copy phage integrase vectors (discussed below).

Electroporation

Enterococcal species are not known to be naturally competent necessitating the use of other methods to introduce foreign DNA into the cell (Fig. 2), with electroporation the method of choice. A common transformation method is the use of stationary phase cells, (which exhibit a 10-fold improvement in transformation efficiency over exponential phase cells), repeated washes in 10% glycerol to remove salts and storage at -70°C (McIntyre and Harlander 1989). They are then electroporated with E. coli isolated plasmid DNA and recovered in nutrient rich media prior to plating on antibiotic selective agar. A variable transformation efficiency was observed, with between 100and 105 per µg of DNA with a range of enterococcal species (5 × 104 per ug for E. faecalis JH2-2) (Friesenegger et al. 1991). However, over half of the enterococcal strains tested were not transformable by this approach. This study tested one E. faecium strain ATCC 9790, which has since been reclassified as E. hirae (Friesenegger et al. 1991, Kristich et al. 2007, Gaechter et al. 2012).

Figure 2.

Figure 2.

Mechanisms of DNA uptake in enterococci. A schematic diagram of DNA uptake in enterococci via transformation of plasmid DNA using electroporation and subsequent conjugation of plasmid DNA to another cell and transduction of phage DNA into the cell. Shown is a type I restriction modification system, a (HsdS) specificity protein and (HsdM) MTase complex (HsdM2HsdS1) has adenine methylated at the HsdS recognition sequences on the host genome to inhibit the generation of the RM complex (HsdM2HsdS1HsdR2). The foreign phage DNA is not methylated and is targeted by the HsdM2HsdS1, which binds the unmethylated recognition site, and then the HsdR2 yielding the RM complex, which then non-specifically induces dsDNA breaks away from the recognition site by collision (Phillips et al. 2019b). Created with BioRender.com.

Further optimisation of the protocol for E. faecalis has included supplementation of the media with glycine to weaken the peptidoglycan in the presence of sucrose to act as an osmotic stabiliser (Shepard and Gilmore 1995). Post growth, all reagents added to the cells are kept on ice (wash buffers, plasmid and cuvettes) as it was shown to increase electroporation efficiency (Shepard and Gilmore 1995). Addition of ice-cold broth post electroporation and incubation on ice to delay the closure of membrane pores improved DNA uptake. Between 150 ng to 1000 ng of plasmid was recommended for transformation as the electroporation efficiencies were shown to decrease over 1 µg, while in S. aureus concentrations up to 10 µg can improve the total number of transformants obtained (Shepard and Gilmore 1995, Monk et al. 2015). The glycine method (Shepard and Gilmore 1995) yielded 1 × 106 per µg of DNA for E. faecalis strain JH2-2, which is a 20-fold improvement over previous efforts (Friesenegger et al. 1991).

For E. faecalis the addition of lysozyme to log phase cells (OD600 0.5–1) to weaken the peptidoglycan improved DNA uptake (Bae et al. 2002). A similar lysozyme or combined with mutanolysin approach was shown to work for E. faecium (Huo et al. 2019, Chen et al. 2021). It has also been observed that a protocol that works for one strain might not be optimal for another, therefore protocols should be tested empirically (Chen et al. 2021) Methods to overcome the poor electroporation efficiency in E. faecium will allow streamlined introduction of plasmid, potentially by using methylated DNA to bypass the RM systems found in E. faecium and E. faecalis. While optimised electroporation parameters are essential to maximise plasmid transfer, such conditions will not impact any endogenous enzymatic restriction of plasmid DNA. Where restriction of DNA is a barrier, application of a plasmid artificial modification approach (as described above for S. aureus (Monk et al. 2015)) through the expression of methylases in E. coli to mimic the adenine and/or cytosine methylation profile of the target strain can help unlock the genetics of hard to transform bacteria (Purdy et al. 2002, O'Connell Motherway et al. 2009, Johnston et al. 2019, Lee et al. 2019, Riley et al. 2019) and could be beneficial for some enterococci.

Transformation via protoplast formation

Protoplast transformation of enterococci has been largely superseded with the advent of electroporation. Reversible generation of E. faecalis protoplasts has been extensively detailed (Wirth et al. 1986). This study optimised the protoplast transformation method using the shuttle vector pAM401 and generated 1 × 106 transformants per µg of plasmid. Cells were grown overnight in Todd-Hewitt broth (THB) with 2%-4% glycine, then sub-cultured in the same media to early log phase, harvested and stored in 50:50 THB: glycerol at -70°C. Thawed cells were enzymatically digested with lysozyme to remove the peptidoglycan for 60–120 minutes in the presence of BSA. Plasmid was added to the protoplasts in buffer containing 40% polyethylene glycol, mixed, incubated on ice and then at 37°C. Cells were recovered in THB with osmotic support at 37°C and plated onto selective media for 2 to 3 days (Wirth et al. 1986). The protoplast transformation method was further optimised by Demuth et al to identify the ideal glycine concentration and lysozyme incubation time (Demuth et al. 1989).

Genetic manipulation techniques in enterococci

The ability to readily and reproducibly introduce to DNA into enterococci opens the door to rapid and sophisticated genetic techniques. Techniques such as Mu transposon mutagenesis, construction of genomic, surface display and CRISPRi libraries, bypassing E. coli for the cloning and expression of toxic genes, phage integrase vectors and rapid allelic exchange all require highly competent cells (greater than 105 CFU transformants)(Poquet et al. 1998, Pajunen et al. 2005, Monk et al. 2008, Monk et al. 2010, Bose et al. 2012, Jiang et al. 2020, Monk and Stinear 2021).

Transposon mutagenesis

Random transposon mutagenesis and construction of transposon mutant libraries is a means to identify genes that are essential or conditionally essential. If a transposon insertion is in an essential gene the cell will die, therefore no transposon mutants will be found. However, insertions into some non-essential genes will lead to a growth-impaired phenotype under certain conditions. The location of the insertion can be determined by DNA sequencing. For example, as a platform for studying the genetic basis of antibiotic resistance, biofilm formation and fitness of E. faecalis in the human GI tract among other phenotypes, an ordered transposon library was created in E. faecalis OG1RF, a plasmid-free strain amenable to genetic manipulation (Dale et al. 2018).

High density transposon mutant libraries have also been generated in two E. faecium isolates using either microarray (isolate E1162) (Zhang et al. 2012) or transposon sequencing (Tn-seq) (isolate E745) to determine insertion sites (Zhang et al. 2017). Tn-seq involves a high-throughput DNA sequencing approach to identify genes that are essential, conditionally essential or exhibit reduced fitness under certain growth conditions e.g. human serum. (van Opijnen and Camilli 2013, Zhang et al. 2017, Cain et al. 2020).

Allelic exchange

Allelic exchange involves the replacement of a target DNA region through homologous recombination, which can introduce a single base change, heterologous gene(s) or gene deletion. Often temperature sensitive plasmids are used to transfer the deletion construct (region of homology surrounding the desired mutation) into the target strain. First, a single crossover event is selected at a non-permissive temperature for plasmid replication (in the presence of antibiotic selection for the plasmid) through integration either upstream or downstream region of the target loci. A shift to a permissive temperature for plasmid replication (without antibiotic selection) leads to plasmid excision from the chromosome and subsequent loss from the cell. If the second crossover event is in the same region as integration, the strain will remain wild type, but if it occurs through the opposite region, it will generate a mutant. Colony PCR can be used to differentiate the wild type from mutant. This process can be very time consuming with the wild-type reversion often dominant (Monk and Foster 2012, Lehman et al. 2016).

The very low transformation efficiency or inability of clinical E. faecium isolates to take up DNA and lack of optimised techniques for allelic exchange has hampered their widespread investigation at a molecular level. To overcome these issues in E. faecium, Nallapareddy et al. developed an improved temperature-sensitive (ts) vector specifically for E. faecium called pTEX5500ts, which is based on the pWV01ts replicon (Nallapareddy et al.2006a). Still issues were encountered with clinical strains of E. faecium, but they were able to transform 5/11 strains tested (Nallapareddy et al.2006a). Using the same ts replicon, pIMAY-Z has been used to manipulate the E. faecium clinical ST796 isolate AUS0233 (Pidot et al. 2018). Even though AUS0233 has intrinsic beta-galactosidase activity, cells containing the plasmid can be discriminated from plasmid free strains on agar containing X-gal, facilitating rapid screening of plasmid excision.

For E. faecium a Cre-lox approach has also been developed, which allows the positive double crossover selection with an antibiotic cassette. A subsequent second round transformation with a temperature sensitive cre-expressing plasmid is used to excise the marker—leaving frt site scars (Zhang et al. 2012). In E. faecalis, a positive selection step to enrich for the second crossover was developed through the toxicity of 5-FU mediated by the upp encoded uracil phosphoribosyl transferase (Kristich et al. 2005). Upp is required for the pyrimidine salvage pathway. However, application of this approach requires the deletion of the chromosomally encoded upp which may subsequently impact on multiple cellular processes. A second approach for positive selection, which has been applied to both E. faecalis and E. faecium is the PheS* system (Kristich et al. 2007b, Panesso et al. 2011). The cell become exquisitely sensitive to the p-chloro-phenylalanine (p-Cl-Phe) in the presence of the mutated phenylalanine tRNA synthetase allele (denoted as PheS*). p-Cl-Phe is added to the broth to stimulate plasmid excision prior to agar selection. Additional benefit of the approach is that the allelic exchange plasmid (based on the suicide vector pORI280 (Leenhouts et al. 1998) containing constitutive LacZ, oriT from pCF10 and PheS*-pCJK47 or pHOU1) can be conjugated from a permissive donor strain CK111 (E. faecalis OG1Sp upp4::P23-repA containing pCF10-101) to either E. faecalis or E. faecium (Kristich et al. 2007b, Panesso et al. 2011).

The Streptococcus pyogenes Cas9 endonuclease has recently been applied to the targeted rapid allelic exchange in E. faecium with positive selection (de Maat et al. 2019), and yielded targeted mutants in three weeks. In an alternate approach, the S. pyogenes Cas9 positive counter-selection has been combined with a single stranded DNA binding protein (RecT recombinase). Cas9 counterselection and enhanced recombination mediated by RecT with either ssDNA or dsDNA template has increased the speed (under two weeks) and precision for the construction of substitutions, insertions and deletions (Chen et al. 2021). However, the size of deletions that can be constructed with ssDNA is currently limited as has been documented in other bacteria Gram-positive bacteria (Oh and van Pijkeren 2014, Penewit et al. 2018). Additionally, dsDNA insertions required a positive selection through the introduction of antibiotic marker. Recently, a new gene editing system for E. faecium was developed using CRISPR-Cas12a (also called Cpf1). The gene-editing Cas12a system only requires a CRISPR RNA (crRNA) for DNA targeting compared to the two RNA molecules required for Cas9 systems. The T-rich (5′-TTTV-3′) protospacer adjacent motif (PAM) sequence for Cas12a is common in low-GC bacteria, such as E. faecium, making this an ideal tool for targeted genetic manipulation. This study validated that the single plasmid Cas12a system could generate clean deletions or insertions within two weeks and is a breakthrough in optimising CRISPR for the genetic manipulation of E. faecium (Chua and Collins 2022). For the Cas9 system, two rounds of cloning (gRNA and homology arms) into a single plasmid are required, while in the Cas12a system this has been streamlined into one round and applies an in vivo cloning approach. Broader application of the Cas9 and Cas12a systems for genome editing will allow the assessment of mutation specificity and the potential for off-target, second site mutations of the respective systems.

These studies highlight the utility of CRISPR-Cas9 and Cas12a machinery as an efficient and timesaving method for genetic manipulation in E. faecium. In situations where gene inactivation is not possible (essential genes) or required (rapid candidate screening) CRISPR interference (CRISPRi) can be applied. CRISPRi uses a catalytically inactive Cas9 (dCas9–unable to cause DNA breaks) that can bind to target DNA and sterically block transcription, which allows for titratable control of target gene expression. A two-plasmid CRISPRi system for E. faecalis has been constructed comprising nisin-inducible (1) dCas9 and (2) guide RNA (Afonina et al. 2020). Analogous to transposon mutagenesis, the production of saturated CRISPRi libraries has been realised through the application of CRISPR adaptation-mediated library manufacturing in vivo. This approach greatly simplifies the construction and reduces the cost of CRISPR guides. These libraries can examine the partial suppression of essential genes and highlight genes involved in resistance pathways that have been overlooked by less complex libraries (Jiang et al. 2020). This approach should also be applicable to enterococci.

Allelic exchange is currently the ‘gold standard’ for genome editing in enterococci and has been used widely for many years (Kristich et al. 2005). However, there is still room for improvement such as the speed for mutant generation e.g. lack of positive selection and poor resolution of blue/white screening due to inherent beta-galactosidase activity (Kristich et al. 2005, Zhang et al. 2012, de Maat et al. 2019). The application of CRISPR technology to genome editing in enterococci is a recent development in the field (de Maat et al. 2019, Chen et al. 2021, Chua and Collins 2022). The advances in the speed of mutant generation (within 2–3 weeks), positive selection, choice of either 5′-NGG-3′ (Cas9) or 5′-TTTV-3′ (Cas12a) PAM, is a major step forward and could revolutionise genetic manipulation of enterococci.

Complementation

Once a gene is deleted/modified a mutant is created that may exhibit a phenotype different from the wild type. Genetic complementation is essential for validation of the genotype/phenotype. Complementation involves the re-introduction of a wild type gene into the mutant to confirm the relationship. The original locus can be restored by ‘reverse’ allelic exchange with the complemented strain marked with a silent restriction site (Karnik et al. 2013). Genes can also be introduced into neutral loci on the chromosome, meaning a site that is not essential and will not generate a fitness cost if altered. Ideally the copy number and expression profile of the complemented gene will mimic the wild-type strain. It is important to whole genome sequence the mutant and complemented strain to ensure that no secondary mutations were generated that could give an altered phenotype. (Karnik et al. 2013, Azizoglu et al. 2014, Arras et al. 2015, Monk et al. 2017, Monk and Stinear 2021).

Shuttle vectors are also a common approach to complement a mutation, as shown for the restoration of cell wall localisation in a complemented E. faecalissrtA mutant (Kristich et al. 2005). While single copy complementation to minimise gene dosage effects is often preferred but can be a more time-consuming option. DebRoy et al. (Debroy et al. 2012) identified a conserved, neutral site (between convergent genes in E. faecalis) for gene insertion through allelic exchange using the pCJK47 counterselection system (described above). This approach has the added advantage of not requiring antibiotic selection to maintain the gene of interest (Debroy et al. 2012).

Another approach are non-replicative phage integrase vectors that lead to single-copy integration at a specific bacterial attachment site (attB) on the chromosome mediated by the integrase/attP on the vector. Examples of these include TP901-1 of Lactococcus lactis and the PSA listeriophage in Listeria monocytogenes(Monk et al. 2008, Koko et al. 2019). One phage integrase vector has been successfully developed for E. faecalis (Yang et al. 2002).

Inducible expression of the complemented gene enables the modulation of gene expression. Systems that have been applied to enterococci include IPTG (Chen et al. 2021), rhamnose (Kristich et al. 2007a) pheromone (Weaver et al. 2017) or agmatine inducible promoters (Linares et al. 2014). For enterococci the nisin-controlled-expression systems (NICE) has also been widely applied (Afonina et al. 2020). The nisRK two component system encoded on the plasmid activates transcription of the PnisA promoter upon sensing the addition of nisin to the media (Bryan et al. 2000). The NICE system is commonly used in enterococci genetics because of the wide dynamic range achieved with a low concentration of nisin that does not inhibit bacterial growth (Eichenbaum et al. 1998). A recent adaption of the NICE system through the construction of a chimeric histidine kinase (NarQ; E. coli extracellular and transmembrane domains/NisK; cytoplasmic region) to change the signal sensed from nisin to nitrate. These modifications yielded a novel nitrate inducible expression system (Mascari et al. 2022). Other well characterised titratable systems such as tetracycline (Bertram et al. 2022) or cumate (Seo and Schmidt-Dannert 2019) could potentially be applied enterococci through the development of enterococcal specific promoters. For the inducible expression systems available for enterococci, there is no standardised comparison of the dynamic range of in the presence of inducer or repression achieved without induction in a single strain background. The application of inducible promoters is ideal for genetic complementation, especially in situations where toxicity is observed or to investigate expression over a dynamic range.

Conclusions and Future Perspectives

In this article we have discussed the current methods and barriers to the genetic manipulation of enterococci. Where RM-systems in enterococci are identified the use of methylation modified plasmids would likely increase transformation efficiency and permit genetic manipulation of diverse clinical E. faecium isolates. Alternatively, the identification of transformable, clinically relevant sequence types (which lack enzymatic and have physical barriers which can be bypassed) would accelerate our ability to conduct genetic studies. Overall, it is known that E. faecalis is easier to transform and manipulate than E. faecium, as the former species has on average higher transformation efficiencies (Ike et al. 1998, Nallapareddy et al. 2006a). Tools and techniques including electroporation, transformation, transposon mutagenesis, and allelic exchange protocols are normally taken from E. faecalis and applied to E. faecium with variable efficiency (Wirth et al. 1986, Cruz-Rodz and Gilmore 1990, Kristich et al. 2005, Kristich et al. 2008). By using the research and techniques reviewed here to create specific protocols for E. faecium it should be possible to apply to powerful modern methods of genetic engineering such as transposon mutagenesis and CRISPR systems, to streamline and optimise mutant generation. The ability to genetically manipulate clinically relevant strains will permit scientists to answer more in-depth questions about the pathogenesis of enterococci, especially with regard to healthcare environments.

Contributor Information

Alexandra L Krause, Department of Microbiology and Immunology, The University of Melbourne at the Peter Doherty Institute for Infection & Immunity, Melbourne, VIC 3000 Australia.

Timothy P Stinear, Department of Microbiology and Immunology, The University of Melbourne at the Peter Doherty Institute for Infection & Immunity, Melbourne, VIC 3000 Australia.

Ian R Monk, Department of Microbiology and Immunology, The University of Melbourne at the Peter Doherty Institute for Infection & Immunity, Melbourne, VIC 3000 Australia.

Funding

T.P.S. is supported by a Leadership Investigator Grant from the NHMRC of Australia (GNT1194325).

Conflicts of interest statement

The authors declare that there are no conflicts of interest.

References

  1. Afonina I, Ong J, Chua Jet al. Multiplex CRISPRi system enables the study of stage-specific biofilm genetic requirements in Enterococcus faecalis. Mbio. 2020;11:e01101–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Agudelo Higuita NI, Huycke MM. Enterococcal Disease, Epidemiology, and Implications for Treatment. Enterococci: From Commensals to Leading Causes of Drug Resistant Infection. In: Gilmore MS, Clewell DB, Ike Y. et al. (eds.). Boston, 2014. [Google Scholar]
  3. Ali L, Blum HE, Sakiotanc T. Detection and characterization of bacterial polysaccharides in drug-resistant enterococci. Glycoconjugate J. 2019;36:429–38. [DOI] [PubMed] [Google Scholar]
  4. Arias CA, Panesso D, Singh KVet al. Cotransfer of antibiotic resistance genes and a hylEfm-containing virulence plasmid in Enterococcus faecium. Antimicrob Agents Chemother. 2009;53:4240–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Arras SD, Chitty JL, Blake KLet al. A genomic safe haven for mutant complementation in Cryptococcus neoformans. PLoS One. 2015;10:e0122916. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Arredondo-Alonso S, Top J, McNally Aet al. Plasmids shaped the recent emergence of the major nosocomial pathogen Enterococcus faecium. Mbio. 2020;11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Atack JM, Guo C, Litfin Tet al. Systematic analysis of REBASE identifies numerous type I restriction-modification systems with duplicated, distinct hsdS specificity genes that can switch system specificity by recombination.mSystems. 2020a;5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Atack JM, Guo C, Yang Let al. DNA sequence repeats identify numerous type I restriction-modification systems that are potential epigenetic regulators controlling phase-variable regulons; phasevarions. FASEB J. 2020b;34:1038–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Atilano ML, Pereira PM, Yates Jet al. Teichoic acids are temporal and spatial regulators of peptidoglycan cross-linking in Staphylococcus aureus. Proc Natl Acad Sci. 2010;107:18991–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Aune TEV, Aachmann FL. Methodologies to increase the transformation efficiencies and the range of bacteria that can be transformed. Appl Microbiol Biotechnol. 2010;85:1301–13. [DOI] [PubMed] [Google Scholar]
  11. Azizoglu RO, Elhanafi D, Kathariou S. Mutant construction and integration vector-mediated gene complementation in Listeria monocytogenes. Methods Mol Biol. 2014;1157:201–11. [DOI] [PubMed] [Google Scholar]
  12. Bae T, Kozlowicz B, Dunny GM. Two targets in pCF10 DNA for PrgX binding: their role in production of qa and prgX mRNA and in regulation of pheromone-inducible conjugation. J Mol Biol. 2002;315:995–1007. [DOI] [PubMed] [Google Scholar]
  13. Belloso Daza MV, Cortimiglia C, Bassi Det al. Genome-based studies indicate that the Enterococcus faecium clade B strains belong to Enterococcus lactis species and lack of the hospital infection associated markers. Int J Syst Evol Microbiol. 2021;71. [DOI] [PubMed] [Google Scholar]
  14. Bertram R, Neumann B, Schuster CF. Status quo of tet regulation in bacteria. Microb Biotechnol. 2022;15:1101–19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Bhattacharjee D, Sorg JA. Factors and conditions that impact electroporation of Clostridioides difficile strains. Msphere. 2020;5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Billington EO, Phang SH, Gregson DBet al. Incidence, risk factors, and outcomes for Enterococcus spp. blood stream infections: a population-based study. Int J Infect Dis. 2014;26:76–82. [DOI] [PubMed] [Google Scholar]
  17. Bose JL, Lehman MK, Fey PDet al. Contribution of the Staphylococcus aureus atl AM and GL murein hydrolase activities in cell division, autolysis, and biofilm formation. PLoS One. 2012;7:e42244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Bourgogne A, Garsin DA, Qin Xet al. Large scale variation in Enterococcus faecalis illustrated by the genome analysis of strain OG1RF. Genome Biol. 2008;9:R110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Brown L, Wolf JM, Prados-Rosales Ret al. Through the wall: extracellular vesicles in Gram-positive bacteria, Mycobacteria and fungi. Nat Rev Microbiol. 2015;13:620–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Brown S, Santa Maria JP Jr, Walker S. Wall teichoic acids of Gram-positive bacteria. Annu Rev Microbiol. 2013;67:313–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Bryan EM, Bae T, Kleerebezem Met al. Improved vectors for nisin-controlled expression in Gram-positive bacteria. Plasmid. 2000;44:183–90. [DOI] [PubMed] [Google Scholar]
  22. Buckley ND, Vadeboncoeur C, LeBlanc DJet al. An effective strategy, applicable to Streptococcus salivarius and related bacteria, to enhance or confer electroporation competence. Appl Environ Microbiol. 1999;65:3800–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Buultjens AH, Lam MM, Ballard Set al. Evolutionary origins of the emergent ST796 clone of vancomycin resistant Enterococcus faecium. PeerJ. 2017;5:e2916. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Byappanahalli MN, Nevers MB, Korajkic Aet al. Enterococci in the environment. Microbiol Mol Biol Rev. 2012;76:685–706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Cain AK, Barquist L, Goodman ALet al. A decade of advances in transposon-insertion sequencing. Nat Rev Genet. 2020;21:526–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Canfield GS, Chatterjee A, Espinosa Jet al. Lytic bacteriophages facilitate antibiotic sensitization of Enterococcus faecium. Antimicrob Agents Chemother. 2021;65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Canfield GS, Duerkop BA. Molecular mechanisms of enterococcal-bacteriophage interactions and implications for human health. Curr Opin Microbiol. 2020;56:38–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Chang JD, Foster EE, Yang Het al. Quantification of the d-Ala-d-Lac-Terminated peptidoglycan structure in vancomycin-resistant Enterococcus faecalis using a combined solid-state nuclear magnetic resonance and mass spectrometry analysis. Biochemistry. 2017;56:612–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Chang JD, Wallace AG, Foster EEet al. Peptidoglycan compositional analysis of Enterococcus faecalis biofilm by stable isotope labeling by amino acids in a bacterial culture. Biochemistry. 2018;57:1274–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Chatterjee A, Johnson CN, Luong Pet al. Bacteriophage resistance alters antibiotic-mediated intestinal expansion of Enterococci. Infect Immun. 2019;87. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Chen V, Griffin ME, Maguin Pet al. RecT recombinase expression enables efficient gene editing in Enterococcus spp. Appl Environ Microbiol. 2021;87:e0084421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Chiang YN, Penades JR, Chen J. Genetic transduction by phages and chromosomal islands: the new and noncanonical. PLoS Pathog. 2019;15:e1007878. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Chua MJ, Collins J. Rapid, efficient, and cost-effective gene editing of Enterococcus faecium with CRISPR-Cas12a. Microbiol Spect. 2022;10:e0242721. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Clark TA, Murray IA, Morgan RDet al. Characterization of DNA methyltransferase specificities using single-molecule, real-time DNA sequencing. Nucleic Acids Res. 2012;40:e29. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Clewell DB, An FY, White BAet al. Sex pheromones and plasmid transfer in Streptococcus faecalis: a pheromone, cAM373, which is also excreted by Staphylococcus aureus. Basic Life Sci. 1985;30:489–503. [DOI] [PubMed] [Google Scholar]
  36. Coombs GW, Pearson JC, Christiansen Ket al. Australian group on antimicrobial resistance enterococcus surveillance programme annual report, 2010. Commun Dis Intell Q Rep. 2013;37:E199–209. [DOI] [PubMed] [Google Scholar]
  37. Cruz-Rodz AL, Gilmore MS. High efficiency introduction of plasmid DNA into glycine treated Enterococcus faecalis by electroporation. Mol Gen Genetics MGG. 1990;224:152–4. [DOI] [PubMed] [Google Scholar]
  38. Dadashi M, Sharifian P, Bostanshirin Net al. The global prevalence of daptomycin, tigecycline, and linezolid-resistant Enterococcus faecalis and Enterococcus faecium strains from human clinical samples: a systematic review and meta-analysis. Frontiers in Medicine. 2021;8:720647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Dale JL, Beckman KB, Willett JLEet al. Comprehensive functional analysis of the Enterococcus faecalis core genome using an ordered, sequence-defined collection of insertional mutations in strain OG1RF. Msystems. 2018;3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. de Maat V, Arredondo-Alonso S, Willems RJLet al. Conditionally essential genes for survival during starvation in Enterococcus faecium E745. BMC Genomics. 2020;21:568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. de Maat V, Stege PB, Dedden Met al. CRISPR-Cas9-mediated genome editing in vancomycin-resistant Enterococcus faecium. FEMS Microbiol Lett. 2019;366. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Debroy S, van der Hoeven R, Singh KVet al. Development of a genomic site for gene integration and expression in Enterococcus faecalis. J Microbiol Methods. 2012;90:1–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Demuth DR, Berthold P, Leboy PSet al. Saliva-mediated aggregation of Enterococcus faecalis transformed with a Streptococcus sanguis gene encoding the SSP-5 surface antigen. Infect Immun. 1989;57:1470–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Di Sante L, Morroni G, Brenciani Aet al. pHTβ-promoted mobilization of non-conjugative resistance plasmids from Enterococcus faecium to Enterococcus faecalis. J Antimicrob Chemother. 2017;72:2447–53. [DOI] [PubMed] [Google Scholar]
  45. Dubin K, Pamer EG, Britton RAet al. Enterococci and their interactions with the intestinal microbiome. Microbiology Spectrum. 2017;5:5.6.01. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Dunny GM, Leonard BA, Hedberg PJ. Pheromone-inducible conjugation in Enterococcus faecalis: interbacterial and host-parasite chemical communication. J Bacteriol. 1995;177:871–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Dunny GM. The peptide pheromone-inducible conjugation system of Enterococcus faecalis plasmid pCF10: cell-cell signalling, gene transfer, complexity and evolution. Philos Trans Royal Soc B: Biolog Sci. 2007;362:1185–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Eichenbaum Z, Federle MJ, Marra Det al. Use of the lactococcal nisA promoter to regulate gene expression in Gram-positive bacteria: comparison of induction level and promoter strength. Appl Environ Microbiol. 1998;64:2763–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Esmail MAM, Abdulghany HM, Khairy RM. Prevalence of multidrug-resistant Enterococcus faecalis in hospital-acquired surgical wound infections and bacteremia: concomitant analysis of antimicrobial resistance genes. Infect Dis: Res Treat. 2019;12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Fiore E, Van Tyne D, Gilmore MS. Chapter 24. Pathogenicity of Enterococci. Gram-Positive Pathogens Third Edition, Wiley, 2019a. [Google Scholar]
  51. Fiore E, Van Tyne D, Gilmore MS. Pathogenicity of enterococci. Microbiol Spect. 2019b;7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Flusberg BA, Webster DR, Lee JHet al. Direct detection of DNA methylation during single-molecule, real-time sequencing. Nat Methods. 2010;7:461–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Fournet-Fayard S, Joly B, Forestier C. Transformation of wild type Klebsiella pneumoniae with plasmid DNA by electroporation. J Microbiol Methods. 1995;24:49–54. [Google Scholar]
  54. Franz CM, Stiles ME, Schleifer KHet al. Enterococci in foods–a conundrum for food safety. Int J Food Microbiol. 2003;88:105–22. [DOI] [PubMed] [Google Scholar]
  55. Friesenegger A, Fiedler S, Devriese LAet al. Genetic transformation of various species of enterococcus by electroporation. FEMS Microbiol Lett. 1991;79:323–7. [DOI] [PubMed] [Google Scholar]
  56. Furmanek-Blaszk B, Sektas M. The SfaNI restriction-modification system from Enterococcus faecalis NEB215 is located on a putative mobile genetic element. FEMS Microbiol Lett. 2015;362:fnv028. [DOI] [PubMed] [Google Scholar]
  57. Gaechter T, Wunderlin C, Schmidheini Tet al. Genome sequence of E nterococcus hirae (Streptococcus faecalis) ATCC 9790, a model organism for the study of ion transport, bioenergetics, and copper homeostasis. J Bacteriol. 2012;194:5126–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Gifford I, Vance S, Nguyen Get al. A stable genetic transformation system and implications of the type IV restriction system in the nitrogen-fixing plant endosymbiont frankia alni ACN14a. Front Microbio. 2019;10:2230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Giltner CL, Nguyen Y, Burrows LL. Type IV pilin proteins: versatile molecular modules. Microbiol Mol Biol Rev. 2012;76:740–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Gonzalez-Ceron G, Miranda-Olivares OJ, Servin-Gonzalez L. Characterization of the methyl-specific restriction system of Streptomyces coelicolor A3(2) and of the role played by laterally acquired nucleases. FEMS Microbiol Lett. 2009;301:35–43. [DOI] [PubMed] [Google Scholar]
  61. Grayson ML, Ballard SA, Gao Wet al. Quantitative efficacy of alcohol-based handrub against vancomycin-resistant enterococci on the hands of human volunteers. Infect Control Hospital Epidemiol. 2012;33:98–100. [DOI] [PubMed] [Google Scholar]
  62. Hammerum AM. Enterococci of animal origin and their significance for public health. Clin Microbiol Infect. 2012;18:619–25. [DOI] [PubMed] [Google Scholar]
  63. Hammes W, Schleifer KH, Kandler O. Mode of action of glycine on the biosynthesis of peptidoglycan. J Bacteriol. 1973;116:1029–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Hancock LE, Murray BE, Sillanpaa J. Enterococcal Cell Wall Components and Structures. In: Gilmore MS, Clewell DB, Ike Y. et al. (eds.). Enterococci: From Commensals to Leading Causes of Drug Resistant Infection. Boston, 2014. [PubMed] [Google Scholar]
  65. Holland LM, Conlon B, O'Gara JP. Mutation of tagO reveals an essential role for wall teichoic acids in Staphylococcus epidermidis biofilm development. Microbiology. 2011;157:408–18. [DOI] [PubMed] [Google Scholar]
  66. Hollenbeck BL, Rice LB. Intrinsic and acquired resistance mechanisms in enterococcus. Virulence. 2012;3:421–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Holo H, Nes IF. High-Frequency transformation, by electroporation, of Lactococcus lactis subsp. cremoris grown with glycine in osmotically stabilized media. Appl Environ Microbiol. 1989;55:3119–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Huang X, Wang J, Li Jet al. Prevalence of phase variable epigenetic invertons among host-associated bacteria. Nucleic Acids Res. 2020;48:11468–85. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Huo W, Adams HM, Trejo Cet al. A type I restriction-modification system associated with Enterococcus faecium subspecies separation. Appl Environ Microbiol. 2019;85. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Huo W, Adams HM, Zhang MQet al. Genome modification in Enterococcus faecalis OG1RF assessed by bisulfite sequencing and single-molecule real-time sequencing. J Bacteriol. 2015;197:1939–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Ike Y, Tanimoto K, Tomita Het al. Efficient transfer of the pheromone-independent Enterococcus faecium plasmid pMG1 (Gmr) (65.1 kilobases) to Enterococcus strains during broth mating. J Bacteriol. 1998;180:4886–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Jacob AE, Hobbs SJ. Conjugal transfer of plasmid-borne multiple antibiotic resistance in Streptococcus faecalis var. zymogenes. J Bacteriol. 1974;117:360–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Jacobs WR Jr Gene transfer in Mycobacterium tuberculosis: shuttle phasmids to enlightenment. Microbiol Spectr. 2014;2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Jain P, Hsu T, Arai Met al. Specialized transduction designed for precise high-throughput unmarked deletions in Mycobacterium tuberculosis. Mbio. 2014;5:e01245–01214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Jeon B, Muraoka W, Scupham Aet al. Roles of lipooligosaccharide and capsular polysaccharide in antimicrobial resistance and natural transformation of C ampylobacter jejuni. J Antimicrob Chemother. 2009;63:462–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Jiang W, Oikonomou P, Tavazoie S. Comprehensive Genome-wide perturbations via CRISPR adaptation reveal complex genetics of antibiotic sensitivity. Cell. 2020;180:1002–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Johnson CN, Sheriff EK, Duerkop BAet al. Let me upgrade you: impact of mobile genetic elements on enterococcal adaptation and evolution. J Bacteriol. 2021;203:e0017721. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Johnston CD, Cotton SL, Rittling SRet al. Systematic evasion of the restriction-modification barrier in bacteria. Proc Natl Acad Sci. 2019;116:11454–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Kajihara T, Nakamura S, Iwanaga Net al. Clinical characteristics and risk factors of enterococcal infections in nagasaki, japan: a retrospective study. BMC Infect Dis. 2015;15:426. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Karnik A, Karnik R, Grefen C. SDM-Assist software to design site-directed mutagenesis primers introducing “silent” restriction sites. BMC Bioinf. 2013;14:105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Kinnear CL, Hansen E, Morley VJet al. Daptomycin treatment impacts resistance in off-target populations of vancomycin-resistant Enterococcus faecium. PLoS Biol. 2020;18:e3000987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Kleiner M, Bushnell B, Sanderson KEet al. Transductomics: sequencing-based detection and analysis of transduced DNA in pure cultures and microbial communities. Microbiome. 2020;8:158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Kohler V, Keller W, Grohmann E. Regulation of Gram-positive conjugation. Frontiers in Microbiology. 2019;10:1134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Koko I, Song AA, Masarudin MJet al. Engineering integrative vectors based on phage site-specific recombination mechanism for Lactococcus lactis. BMC Biotech. 2019;19:82. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Komiyama EY, Lepesqueur LS, Yassuda CGet al. Enterococcus species in the oral cavity: prevalence, virulence factors and antimicrobial susceptibility. PLoS One. 2016;11:e0163001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Krawczyk B, Wityk B, Gałęcka Met al. The Many Faces of Enterococcus spp.-Commensal, Probiotic and Opportunistic Pathogen. Microorganisms. 2021;9:1900. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Kreiswirth BN, Lofdahl S, Betley MJet al. The toxic shock syndrome exotoxin structural gene is not detectably transmitted by a prophage. Nature. 1983;305:709–12. [DOI] [PubMed] [Google Scholar]
  88. Kristich CJ, Chandler JR, Dunny GM. Development of a host-genotype-independent counterselectable marker and a high-frequency conjugative delivery system and their use in genetic analysis of Enterococcus faecalis. Plasmid. 2007b;57:131–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Kristich CJ, Manias DA, Dunny GM. Development of a method for markerless genetic exchange in Enterococcus faecalis and its use in construction of a srtA mutant. Appl Environ Microbiol. 2005;71:5837–49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Kristich CJ, Nguyen VT, Le Tet al. Development and use of an efficient system for random mariner transposon mutagenesis to identify novel genetic determinants of biofilm formation in the core Enterococcus faecalis genome. Appl Environ Microbiol. 2008;74:3377–86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Kristich CJ, Rice LB, Arias CA. Enterococcal Infection-Treatment and Antibiotic Resistance. In: Gilmore MS, Clewell DB, Ike Y. et al. (eds.) Enterococci: From Commensals to Leading Causes of Drug Resistant Infection. Boston, 2014. [PubMed] [Google Scholar]
  92. Kristich CJ, Wells CL, Dunny GM. A eukaryotic-type Ser/Thr kinase in Enterococcus faecalis mediates antimicrobial resistance and intestinal persistence. Proc Natl Acad Sci. 2007a;104:3508–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Leblanc DJ, Cohen L, Jensen L. Transformation of group F Streptococci by plasmid DNA. J Gen Microbiol. 1978;106:49–54. [DOI] [PubMed] [Google Scholar]
  94. Lebreton F, van Schaik W, McGuire AMet al. Emergence of epidemic multidrug-resistant Enterococcus faecium from animal and commensal strains. Mbio. 2013;4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Lebreton F, Willems RJL, Gilmore MS. Enterococcus Diversity, Origins in Nature, and Gut Colonization. In: Gilmore MS, Clewell DB, Ike Y, et al. (eds.). Enterococci: From Commensals to Leading Causes of Drug Resistant Infection. Boston, 2014. [PubMed] [Google Scholar]
  96. Lee JYH, Carter GP, Pidot SJet al. Mining the methylome reveals extensive diversity in Staphylococcus epidermidis restriction modification. Mbio. 2019;10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Leenhouts K, Venema G, Kok J. A lactococcal pWV01-based integration toolbox for bacteria. Methods Cell Sci. 1998;20:35–50. [Google Scholar]
  98. Lehman MK, Bose JL, Bayles KW. Allelic exchange. Methods Mol Biol. 2016;1373:89–96. [DOI] [PubMed] [Google Scholar]
  99. Li L, Higgs C, Turner AMet al. Daptomycin resistance occurs predominantly in vanA-Type vancomycin-resistant Enterococcus faecium in australasia and is associated with heterogeneous and novel mutations. Frontiers in Microbiology. 2021;12:749935. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Linares DM, Perez M, Ladero Vet al. An agmatine-inducible system for the expression of recombinant proteins in Enterococcus faecalis. Microb Cell Fact. 2014;13:169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Loenen WA, Dryden DT, Raleigh EAet al. Highlights of the DNA cutters: a short history of the restriction enzymes. Nucleic Acids Res. 2014;42:3–19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Malisova L, Jakubu V, Pomorska Ket al. Spread of linezolid-resistant Enterococcus spp. in human clinical isolates in the Czech Republic. Antibiotics (Basel). 2021;10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Manso AS, Chai MH, Atack JMet al. A random six-phase switch regulates pneumococcal virulence via global epigenetic changes. Nat Commun. 2014;5:5055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Manson JM, Smith JM, Cook GM. Persistence of vancomycin-resistant enterococci in New Zealand broilers after discontinuation of avoparcin use. Appl Environ Microbiol. 2004;70:5764–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Marcinek H, Wirth R, Muscholl-Silberhorn Aet al. Enterococcus faecalis gene transfer under natural conditions in municipal sewage water treatment plants. Appl Environ Microbiol. 1998;64:626–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Mascari CA, Djoric D, Little JLet al. Use of an interspecies chimeric receptor for inducible gene expression reveals that metabolic flux through the peptidoglycan biosynthesis pathway is an important driver of cephalosporin resistance in Enterococcus faecalis. J Bacteriol. 2022;204:e0060221. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Mazaheri Nezhad Fard R, Barton MD, Heuzenroeder MW. Bacteriophage-mediated transduction of antibiotic resistance in enterococci. Lett Appl Microbiol. 2011;52:559–64. [DOI] [PubMed] [Google Scholar]
  108. McIntyre DA, Harlander SK. Improved electroporation efficiency of intact Lactococcus lactis subsp. lactis cells grown in defined media. Appl Environ Microbiol. 1989;55:2621–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. McMahon SA, Roberts GA, Johnson KAet al. Extensive DNA mimicry by the ArdA anti-restriction protein and its role in the spread of antibiotic resistance. Nucleic Acids Res. 2009;37:4887–97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Mell JC, Redfield RJ. Natural competence and the evolution of DNA uptake specificity. J Bacteriol. 2014;196:1471–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Mercuro NJ, Davis SL, Zervos MJet al. Combatting resistant enterococcal infections: a pharmacotherapy review. Expert Opin Pharmacother. 2018;19:979–92. [DOI] [PubMed] [Google Scholar]
  112. Mlaga KD, Garcia V, Colson Pet al. Extensive comparative genomic analysis of Enterococcus faecalis and Enteroc occus faecium reveals a direct association between the absence of CRISPR-Cas systems, the presence of anti-endonuclease (ardA) and the acquisition of vancomycin resistance in E . faecium. Microorganisms. 2021;9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Monk IR, Casey PG, Hill Cet al. Directed evolution and targeted mutagenesis to murinize Li steria monocytogenes internalin a for enhanced infectivity in the murine oral infection model. BMC Microbiol. 2010;10:318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Monk IR, Foster TJ. Genetic manipulation of Staphylococci-breaking through the barrier. Front Cell Infect Microbiol. 2012;2:49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Monk IR, Gahan CG, Hill C. Tools for functional postgenomic analysis of L isteria monocytogenes. Appl Environ Microbiol. 2008;74:3921–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Monk IR, Howden BP, Seemann Tet al. Correspondence: spontaneous secondary mutations confound analysis of the essential two-component system WalKR in Staphylococcus aureus. Nat Commun. 2017;8:14403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Monk IR, Shah IM, Xu Met al. Transforming the untransformable: application of direct transformation to manipulate genetically Staphylococcus aureus and Staphylococcus epidermidis. Mbio. 2012;3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Monk IR, Stinear TP. From cloning to mutant in 5 days: rapid allelic exchange in Staphylococcus aureus. Access Microbiology. 2021;3:000193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Monk IR, Tree JJ, Howden BPet al. Complete bypass of restriction systems for major S taphylococcus aureus lineages. Mbio. 2015;6:e00308–00315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Murray NE. Type I restriction systems: sophisticated molecular machines (a legacy of Bertani and Weigle). Microbiol Mol Biol Rev. 2000;64:412–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Muschiol S, Balaban M, Normark Set al. Bioessays. 2015;37:426–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Nallapareddy SR, Singh KV, Murray BE. Construction of improved temperature-sensitive and mobilizable vectors and their use for constructing mutations in the adhesin-encoding acm gene of poorly transformable clinical Enterococcus faecium strains. Appl Environ Microbiol. 2006a;72:334–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Nallapareddy SR, Singh KV, Sillanpaa Jet al. Endocarditis and biofilm-associated pili of Enterococcus faecalis. J Clin Invest. 2006b;116:2799–807. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Nidhi S, Anand U, Oleksak Pet al. Novel CRISPR-Cas systems: an updated review of the current achievements, applications, and future research perspectives. Int J Mol Sci. 2021;22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Nuesch-Inderbinen M, Raschle S, Stevens MJAet al. Linezolid-resistant Enterococcus faecalis ST16 harbouring optrA on a Tn6674-like element isolated from surface water. J Global Antimicro Resist. 2021;25:89–92. [DOI] [PubMed] [Google Scholar]
  126. O'Connell Motherway M, O'Driscoll J, Fitzgerald GFet al. Overcoming the restriction barrier to plasmid transformation and targeted mutagenesis in Bifidobacterium breve UCC2003. Microb Biotechnol. 2009;2:321–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Oh JH, van Pijkeren JP. CRISPR-Cas9-assisted recombineering in Lactobacillus reuteri. Nucleic Acids Res. 2014;42:e131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Okhapkina SS, Netesova NA, Golikova LNet al. Comparison of the homologous SfeI and LlaBI restriction-modification systems. Mol Biol. 2002;36:432–7. [PubMed] [Google Scholar]
  129. Oliveira PH, Touchon M, Rocha EP. The interplay of restriction-modification systems with mobile genetic elements and their prokaryotic hosts. Nucleic Acids Res. 2014;42:10618–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Pajunen MI, Pulliainen AT, Finne Jet al. Generation of transposon insertion mutant libraries for Gram-positive bacteria by electroporation of phage mu DNA transposition complexes. Microbiology. 2005;151:1209–18. [DOI] [PubMed] [Google Scholar]
  131. Palmer KL, Godfrey P, Griggs Aet al. Comparative genomics of enterococci: variation in Enterococcus faecalis, clade structure in E . faecium, and defining characteristics of E . gallinarum and E . casseliflavus. Mbio. 2012;3:e00318–00311. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Panesso D, Montealegre MC, Rincon Set al. The hylEfm gene in pHylEfm of Enterococcus faecium is not required in pathogenesis of murine peritonitis. BMC Microbiol. 2011;11:20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Paulsen IT, Banerjei L, Myers GSet al. Role of mobile DNA in the evolution of vancomycin-resistant Enterococcus faecalis. Science. 2003;299:2071–4. [DOI] [PubMed] [Google Scholar]
  134. Payne LJ, Todeschini TC, Wu Yet al. Identification and classification of antiviral defence systems in bacteria and archaea with PADLOC reveals new system types. Nucleic Acids Res. 2021;49:10868–78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Penewit K, Holmes EA, McLean Ket al. Efficient and scalable precision genome editing in S taphylococcus aureus through conditional recombineering and CRISPR/Cas9-Mediated counterselection. Mbio. 2018;9:e00067–00018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Perry JA, Wright GD. The antibiotic resistance “mobilome”: searching for the link between environment and clinic. Front Microbiol. 2013;4:138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Phillips ZN, Husna AU, Jennings MPet al. Phasevarions of bacterial pathogens - phase-variable epigenetic regulators evolving from restriction-modification systems. Microbiology. 2019b;165:917–28. [DOI] [PubMed] [Google Scholar]
  138. Phillips ZN, Tram G, Seib KLet al. Phase-variable bacterial loci: how bacteria gamble to maximise fitness in changing environments. Biochem Soc Trans. 2019a;47:1131–41. [DOI] [PubMed] [Google Scholar]
  139. Picton DM, Luyten YA, Morgan RDet al. The phage defence island of a multidrug resistant plasmid uses both BREX and type IV restriction for complementary protection from viruses. Nucleic Acids Res. 2021;49:11257–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Pidgeon SE, Apostolos AJ, Nelson JMet al. L,D-Transpeptidase specific probe reveals spatial activity of peptidoglycan cross-linking. ACS Chem Biol. 2019;14:2185–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Pidot SJ, Gao W, Buultjens AHet al. Increasing tolerance of hospital Enterococcus faecium to handwash alcohols. Sci Transl Med. 2018;10. [DOI] [PubMed] [Google Scholar]
  142. Piepenbrink KH. DNA uptake by type IV filaments. Front Mol Biosci. 2019;6:1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Pingoud A, Wilson GG, Wende W. Type II restriction endonucleases–a historical perspective and more. Nucleic Acids Res. 2014;42:7489–527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Poquet I, Ehrlich SD, Gruss A. An export-specific reporter designed for Gram-positive bacteria: application to L actococcus lactis. J Bacteriol. 1998;180:1904–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Purdy D, O'Keeffe TA, Elmore Met al. Conjugative transfer of clostridial shuttle vectors from Escherichia coli to clostridium difficile through circumvention of the restriction barrier. Mol Microbiol. 2002;46:439–52. [DOI] [PubMed] [Google Scholar]
  146. Ramos Y, Sansone S, Morales DK. Sugarcoating it: enterococcal polysaccharides as key modulators of host-pathogen interactions. PLoS Pathog. 2021;17:e1009822. [DOI] [PMC free article] [PubMed] [Google Scholar]
  147. Raven KE, Reuter S, Reynolds Ret al. A decade of genomic history for healthcare-associated Enterococcus faecium in the United Kingdom and Ireland. Genome Res. 2016;26:1388–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Rédei GP. Specialized Transduction. Encyclopedia of Genetics, Genomics, Proteomics and Informatics, p.^pp.1849–. Springer Netherlands: Dordrecht, 2008. [Google Scholar]
  149. Reyes Ruiz LM, Williams CL, Tamayo R. Enhancing bacterial survival through phenotypic heterogeneity. PLoS Pathog. 2020;16:e1008439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Riley LA, Ji L, Schmitz RJet al. Rational development of transformation in Clostridium thermocellum ATCC 27405 via complete methylome analysis and evasion of native restriction-modification systems. J Ind Microbiol Biotechnol. 2019;46:1435–43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Roberts RJ. How restriction enzymes became the workhorses of molecular biology. Proc Natl Acad Sci. 2005;102:5905–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Schleifer KH, Kilpper-Bälz R. Molecular and chemotaxonomic approaches to the classification of streptococci, enterococci and lactococci: a review. Syst Appl Microbiol. 1987;10:1–19. [Google Scholar]
  153. Seo SO, Schmidt-Dannert C. Development of a synthetic cumate-inducible gene expression system for bacillus. Appl Microbiol Biotechnol. 2019;103:303–13. [DOI] [PubMed] [Google Scholar]
  154. Shan X, Yang M, Wang Net al. Plasmid fusion and recombination events that occurred during conjugation of poxtA-Carrying plasmids in enterococci. Microbiol Spect. 2022;10:e0150521. [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Shepard BD, Gilmore MS. Electroporation and efficient transformation of Enterococcus faecalis grown in high concentrations of glycine. Methods Mol Biol. 1995;47:217–26. [DOI] [PubMed] [Google Scholar]
  156. Shockman GD, Barrett JF. Structure, function, and assembly of cell walls of Gram-positive bacteria. Annu Rev Microbiol. 1983;37:501–27. [DOI] [PubMed] [Google Scholar]
  157. Simonsen L, Gordon DM, Stewart FMet al. Estimating the rate of plasmid transfer: an end-point method. J Gen Microbiol. 1990;136:2319–25. [DOI] [PubMed] [Google Scholar]
  158. Sitaraman R. The role of DNA restriction-modification systems in the biology of Bacillus anthracis. Front Microbiol. 2016;7:11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Smith RE, Salamaga B, Szkuta Pet al. Decoration of the enterococcal polysaccharide antigen EPA is essential for virulence, cell surface charge and interaction with effectors of the innate immune system. PLoS Pathog. 2019;15:e1007730. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Staddon JH, Bryan EM, Manias DAet al. Genetic characterization of the conjugative DNA processing system of enterococcal plasmid pCF10. Plasmid. 2006;56:102–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Sterling AJ, Snelling WJ, Naughton PJet al. Competent but complex communication: the phenomena of pheromone-responsive plasmids. PLoS Pathog. 2020;16:e1008310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Strand TA, Lale R, Degnes KFet al. A new and improved host-independent plasmid system for RK2-based conjugal transfer. PLoS One. 2014;9:e90372. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Sung JM, Lindsay JA. Staphylococcus aureus strains that are hypersusceptible to resistance gene transfer from enterococci. Antimicrob Agents Chemother. 2007;51:2189–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Theilacker C, Holst O, Lindner Bet al. The structure of the wall teichoic acid isolated from Enterococcus faecalis strain 12030. Carbohydr Res. 2012;354:106–9. [DOI] [PubMed] [Google Scholar]
  165. Thurlow LR, Thomas VC, Fleming SDet al. Enterococcus faecalis capsular polysaccharide serotypes C and D and their contributions to host innate immune evasion. Infect Immun. 2009;77:5551–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  166. Timmermans M, Bogaerts B, Vanneste Ket al. Large diversity of linezolid-resistant isolates discovered in food-producing animals through linezolid selective monitoring in Belgium in 2019. J Antimicrob Chemother. 2021;77:49–57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Turgeon N, Laflamme C, Ho Jet al. Elaboration of an electroporation protocol for B acillus cereus ATCC 14579. J Microbiol Methods. 2006;67:543–8. [DOI] [PubMed] [Google Scholar]
  168. van Hal SJ, Willems RJL, Gouliouris Tet al. The interplay between community and hospital Enterococcus faecium clones within health-care settings: a genomic analysis. The Lancet Microbe. 2022;3:e133–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. van Opijnen T, Camilli A. Transposon insertion sequencing: a new tool for systems-level analysis of microorganisms. Nat Rev Microbiol. 2013;11:435–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. van Pijkeren JP, Morrissey D, Monk IRet al. A novel Li steria monocytogenes-based DNA delivery system for cancer gene therapy. Hum Gene Ther. 2010;21:405–16. [DOI] [PubMed] [Google Scholar]
  171. Wang S, Guo Y, Lv Jet al. Characteristic of Enterococcus faecium clinical isolates with quinupristin/dalfopristin resistance in china. BMC Microbiol. 2016;16:246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Weaver KE, Chen Y, Miiller EMet al. Examination of Enterococcus faecalis toxin-antitoxin system toxin fst function utilizing a pheromone-inducible expression vector with tight repression and broad dynamic range. J Bacteriol. 2017;199:e00065–00017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Winstel V, Liang C, Sanchez-Carballo Pet al. Wall teichoic acid structure governs horizontal gene transfer between major bacterial pathogens. Nat Commun. 2013;4:2345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Wirth R, An FY, Clewell DB. Highly efficient protoplast transformation system for S treptococcus faecalis and a new E scherichia coli-S. faecalis shuttle vector. J Bacteriol. 1986;165:831–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Woods C, Humphreys CM, Rodrigues RMet al. A novel conjugal donor strain for improved DNA transfer into C lostridium spp. Anaerobe. 2019;59:184–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Xu SY, Corvaglia AR, Chan SHet al. A type IV modification-dependent restriction enzyme SauUSI from S taphylococcus aureus subsp. aureus USA300. Nucleic Acids Res. 2011;39:5597–610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Yang H, Singh M, Kim SJet al. Characterization of the tertiary structure of the peptidoglycan of Enterococcus faecalis. Biochimica Et Biophysica Acta (BBA) - Biomembranes. 2017;1859:2171–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Yang HY, Kim YW, Chang HI. Construction of an integration-proficient vector based on the site-specific recombination mechanism of enterococcal temperate phage phiFC1. J Bacteriol. 2002;184:1859–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Yasmin A, Kenny JG, Shankar Jet al. Comparative genomics and transduction potential of Enterococcus faecalis temperate bacteriophages. J Bacteriol. 2010;192:1122–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Zavil'gel'skii GB, Rastorguev SM. Antirestriction proteins ArdA and Ocr as effective inhibitors of the type I restriction-modification enzymes. Mol Biol (Mosk). 2009;43:264–73. [PubMed] [Google Scholar]
  181. Zhang X, de Maat V, Guzman Prieto AMet al. RNA-seq and Tn-seq reveal fitness determinants of vancomycin-resistant Enterococcus faecium during growth in human serum. BMC Genomics. 2017;18:893. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Zhang X, Paganelli FL, Bierschenk Det al. Genome-wide identification of ampicillin resistance determinants in Enterococcus faecium. PLos Genet. 2012;8:e1002804. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Zinder ND, Lederberg J. Genetic exchange in salmonella. J Bacteriol. 1952;64:679–99. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from FEMS Microbiology Reviews are provided here courtesy of Oxford University Press

RESOURCES