Next Article in Journal
Surgical Antimicrobial Prophylaxis in Neonates and Children Undergoing Neurosurgery: A RAND/UCLA Appropriateness Method Consensus Study
Next Article in Special Issue
Nisin Mutant Prevention Concentration and the Role of Subinhibitory Concentrations on Resistance Development by Diabetic Foot Staphylococci
Previous Article in Journal
Synergistic Role of Plant Extracts and Essential Oils against Multidrug Resistance and Gram-Negative Bacterial Strains Producing Extended-Spectrum β-Lactamases
Previous Article in Special Issue
Genomic Characterization of Clinical Acinetobacter baumannii Isolates Obtained from COVID-19 Patients in Russia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Enterococcus Virulence and Resistant Traits Associated with Its Permanence in the Hospital Environment

by
Catarina Geraldes
1,2,
Luís Tavares
1,2,
Solange Gil
1,2,3 and
Manuela Oliveira
1,2,*
1
Centre for Interdisciplinary Research in Animal Health (CIISA), Faculty of Veterinary Medicine, University of Lisbon, Av. da Universidade Técnica de Lisboa, 1300-477 Lisbon, Portugal
2
Laboratório Associado para Ciência Animal e Veterinária (AL4AnimalS), 1300-477 Lisbon, Portugal
3
Biological Isolation and Containment Unit (BICU), Veterinary Hospital, Faculty of Veterinary Medicine, University of Lisbon, Av. Universidade Técnica, 1300-477 Lisbon, Portugal
*
Author to whom correspondence should be addressed.
Antibiotics 2022, 11(7), 857; https://doi.org/10.3390/antibiotics11070857
Submission received: 9 June 2022 / Revised: 23 June 2022 / Accepted: 24 June 2022 / Published: 26 June 2022
(This article belongs to the Special Issue Antimicrobial Resistance and Virulence – 3rd Volume)

Abstract

:
Enterococcus are opportunistic pathogens that have been gaining importance in the clinical setting, especially in terms of hospital-acquired infections. This problem has mainly been associated with the fact that these bacteria are able to present intrinsic and extrinsic resistance to different classes of antibiotics, with a great deal of importance being attributed to vancomycin-resistant enterococci. However, other aspects, such as the expression of different virulence factors including biofilm-forming ability, and its capacity of trading genetic information, makes this bacterial genus more capable of surviving harsh environmental conditions. All these characteristics, associated with some reports of decreased susceptibility to some biocides, all described in this literary review, allow enterococci to present a longer survival ability in the hospital environment, consequently giving them more opportunities to disseminate in these settings and be responsible for difficult-to-treat infections.

1. Introduction

Enterococcus spp. are ubiquitous Gram-positive and facultative anaerobic bacteria, commensal of the intestinal tract of humans, as well as other mammals [1,2,3]. These microorganisms are largely characterized by their ability to tolerate high concentrations of salt (6.5% NaCl), but also a wide range of temperature (from 10 °C to 40 °C) and pH (from 4.4 to 9.6) values [1,2]. They are also able to hydrolyze esculin in the presence of high quantities of bile salts (40%) [2].
Initially portrayed as organisms of little clinical importance, enterococci, particularly Enterococcus faecalis and Enterococcus faecium, have been progressively associated with an increasing number of hospital-acquired infections (HAIs) in both human and veterinary medicine [3]. In fact, enterococci account for 6.1–17.5% of all isolates retrieved between 2010 and 2020 from European patients with these types of infections in human medicine [4]. Apart from these two species, other enterococcal species can also be isolated, such as Enterococcus hirae, Enterococcus durans, Enterococcus gallinarium and Enterococcus casseliflavus, not only in veterinary medicine [3], but also in human medicine [5,6,7], although they are not as commonly associated with HAIs and thus are less commonly studied. Enterococci are associated with a wide range of infections including urinary tract infections, bacteremia, endocarditis, wound infections (burn wounds or surgical incisions), abdomen and biliary tract infections, and infection of catheters and medical implants [8].
This infection-inducing capacity becomes especially critical when its degree of antibiotic resistance is considered. The intrinsic and extrinsic resistance traits associated with this genus allow it to be resistant to several antibiotics, including β-lactams, aminoglycosides and glycopeptides, rendering it difficult to combat these infections [1,3,4,8,9]. In fact, vancomycin-resistant Enterococcus faecium is considered by the World Health Organization (WHO) [10] as a high priority pathogen for which new antimicrobial therapies are needed.
This overly known resistance to antibiotics is not the only trait that makes Enterococcus a threat to human and animal health. They also present a great capacity of persisting in the environment [11,12], which might be associated with some reports of decreased susceptibility to certain biocides [13,14,15], especially in the presence of organic matter [16,17]. They also have a formidable biofilm-forming capacity [8], and they are known for their genome’s plasticity, which allows them to easily acquire, conserve and disseminate genetic traits among not only enterococci, but also other Gram-positive bacteria [8,18,19,20,21].
Bearing this in mind, this article presents a review of all the attributes that can justify the extended permanence of this genus in the hospital environment, which consequently leads them to be one of the most important bacteria in terms of HAIs.

2. Genetic Organization

It is said that enterococci capacity to acquire new genetic material is one of the most important traits that allows them to adapt to different environments [22]. This becomes even more evident when we consider that the complete genome sequence of the first vancomycin-resistant E. faecalis (V583) reported in the United States revealed that more than a quarter of its genome consisted of acquired DNA [23].
This adaptability capacity to different environments seems to have divided the E. faecium population into two different clades: clade A, more adapted to the hospital environment, and clade B, considered the community adapted clade [24,25,26,27,28,29,30]. The genome of clade A isolates is usually larger and contains a higher quantity of exogenous genetic material frequently associated to antibiotic resistance and the carbohydrate metabolism [22,25,31], and has been mostly associated to clonal complex (CC) 17 [24,32]. However, this division has been contested since some isolates do not group genetically in this clade [26,29]. Some authors also defend that clade A can be divided into two different sub-clades: clade A1, comprised of human clinical strains, and clade A2, usually an animal-associated group [28,32]; however, this separation branch has also been contested since recent studies found no proof for the existence of this sub-division [29,30].
Regarding E. faecalis, no such genetic division has been made since no correlation has been found between clonal structure and isolate origin [27,33,34]. However, some clonal complexes such as CC2, CC9 and CC87, termed high-risk enterococcal clonal complexes (HiRECCs) due to their multi-drug resistant (MDR) profile, have been more associated to the hospital environment and nosocomial infections [24,33,35].

3. Virulence

In the last few years, vancomycin-resistant Enterococcus faecium have been a rising cause of concern in terms of enterococcal infections. However, when it comes to virulence factors, E. faecalis has the leading role, which explains why it is still considered the primary species in terms of nosocomial infections. Although they are not a main issue in the Enterococcus genus, they still represent an advantageous feature in terms of environmental survival since some of these factors can be associated to biofilm formation and adherence to a variety of surfaces. Virulence factors can be divided into two distinct groups: those that are secreted, and those that are present in the surface of the bacterial cell [32]. In this article, we present the most important virulence factors included in both of these categories.

3.1. Secreted Virulence Factors

One of the first virulence factors described in enterococci was cytolysin, encoded by the genes cylLL and cylLS [36], and named due to its dual action, since it presents both a bactericidal and cytolytic activity [37,38,39]. Cytolysin seems to present some activity against red and white blood cells from mice [40], and also seems to contribute to endophthalmitis [41] and endocarditis [42]. Although a study by Jett et al. in 1992 [41] on a rabbit model indicated a possible association between cytolysin-producing enterococci and endophthalmitis, more recent studies do not seem to find any connection between the two [43,44].
Exclusively present in E. faecium isolates [31], secreted antigen A (SagA) is a stress-related protein [45] that supposedly also plays a role in cell growth, conceivably due to interactions with the cell wall metabolism [46]. It can additionally bind to a series of extracellular matrix proteins [47]. On the other hand, this protein has also been correlated to a better functioning intestinal barrier and enhanced tolerance against other enteric pathogens such as Salmonella Typhimurium and Clostridium difficile, possibly due to activation of the host innate immune system [48,49,50]. SagA has also been associated with biofilm formation, but only in relation to enterococci belonging to clade A1 [50].
Gelatinase, encoded by the gelE gene, is a matrix metalloproteinase, best studied in E. faecalis [8,32]. This protein is capable of hydrolyzing gelatin and collagen and not only seems to be able to interfere with complement-mediated immunity [51], but also seems to be an aid in the development of infectious endocarditis by E. faecalis [52]. GelE is co-transcribed along with SprE, a serine protease, by the FSR system [53,54], which also appears to regulate other types of proteins, some of them associated with biofilm formation [55]. GelE seems to reduce the incidence of Ace, another virulence factor present at the cell surface, possibly due to a GelE-dependent cleavage of the Ace protein, which reduces the enterococci ability to bind to collagen [56].
The hylEfm genes encode a glucosyl hydrolase, and is predominantly present in clinical isolates, especially E. faecium CC17 [57,58]. Although, to the best of our knowledge, there are not many studies that indicate the potential of this protein as a virulence factor, it has been suggested that it could lead to the co-transference and dissemination of antibiotic resistance genes [59].

3.2. Cell Surface Virulence Factors

Cell surface components have been found to be important in a series of different bacterial defense mechanisms, including biofilm formation and protection against the host immune system [60,61].
In enterococci, these surface proteins are usually of the LPxTG-type (leucine, proline, X (any amino acid), threonine, and glycine), which include pili and microbial surface components recognizing adhesive matrix molecules (MSCRAMMs) [62].
The first ever described cell surface group of proteins included the aggregation substances (also known as AS or Agg), which depict a group of three adhesins: Asp1, Asc10 (sometimes also called PrgB) and Asa1. These proteins are encoded by three distinct conjugative plasmids, pPD1, pCF10 and pAD1, respectively, and are very similar in amino acid sequence [62]. These AS have been considered to be responsible for enterococcal adhesion to intestinal [63,64,65] and renal tubular cells [66]. As such, this is possibly associated with their ability to prompt systemic infections [62], since adhesion is usually the first step necessary for infection. This takes an even greater importance knowing that Asa1 adheres to extracellular matrix proteins, which also facilitates infection development [67]. Furthermore, AS have also been correlated to adhesion to immune system cells (especially Asa1 and Asc10) [68,69], and phagocytosis survival [69,70], as well as vanA (an operon associated to vancomycin resistance) co-transference in E. faecalis [71]. Asc10, which is encoded by the prgB gene, has also been associated to a higher virulence in infective endocarditis [72] and biofilm formation [73].
There are three main MSCRAMMs, which are a subfamily of bacterial adhesins that recognize and bind to extracellular matrix elements, described in enterococci: Ace (adhesion of collagen, from E. faecalis) [74], Acm (adhesion of collagen, from E. faecium) [75] and Scm (second collagen adhesion, from E. faecium) [76].
Ace was the first MSCRAMM identified in enterococci [74], followed by Acm [75], and lastly by Scm [76]. Although they all promote adhesion to collagen, Ace has also been characterized by its capacity to adhere to other components such as laminin [77] and dentin [78], while Scm also binds to fibrinogen [76]. Both Ace and Acm seem to play a role in the development of infectious endocarditis caused by enterococci [79,80,81]. Although all three are present in clinical and non-clinical isolates, Acm seems to be predominantly distributed among E. faecium clinical isolates, with a higher incidence among CC17, which could be one of the reasons this complex gained such importance as a nosocomial pathogen [82].
Pili are another relevant virulence factor that consist of multimeric fibres composed of pilin subunits that extend as long filaments from the cell surfaces. In enterococci, these fibres are encoded by an operon consisting of a collection of different genes, commonly called pili gene clusters (PGC).
The E. faecalis genome contains two distinct PGC: EBP (endocarditis and biofilm-associated pilus), composed of three distinct genes, ebpA, ebpB and ebpC; and BEE (biofilm enhancer in Enterococcus), composed of three pili-enconding genes, bee-1, bee-2 and bee-3, and two sortase-like enzyme-encoding genes, srt-1 and srt-2 [83]. While EBP seems to be ubiquitously distributed among E. faecalis [84], the BEE PGC is much less frequently found [83]. Both these clusters seem to be important in biofilm formation [83,85].
On the other hand, E. faecium is considered to have four different gene clusters, which were identified as PGC1 to 4 [62,76,86]. PGC3, known as EMP (previously identified as EBPfm) is composed by three different genes, empA, empB and empC [87,88], and has been associated with biofilm formation and adherence to EMCs [88], as well as colonization of both the kidneys and the bladder in an experimental urinary tract infection model [87].
Esp is another cell surface protein, present in both E. faecium and E. faecalis, which seems to be predominantly present in clinical isolates [89,90]. This protein has been proven to be important in biofilm formation [91,92,93], possibly through an amyloid-based mechanism [94], but has also been implicated in infectious endocarditis [95] and urinary tract infection development [96]. Furthermore, a study published in 2009 by Meredith et al. [97] also indicated a possible relation between the presence of this protein and E. faecium’s susceptibility to β-lactams, which has also been corroborated by a more recent study [98].
Other cell surface proteins acting as potential virulence factors have also been described, such as PrpA [99], that binds to fibrinogen, fibronectin and platelets, and SgrA and EcbA, two adhesins that seem to be predominantly present in Enterococcus clinical isolates [100].
The most important virulence factors presented by the Enterococcus genus can be seen summarized in Figure 1.

4. Antibiotic Resistance

4.1. β-Lactam Resistance

Enterococcus present a variable intrinsic resistance to the different classes of β-lactams; they usually present only a reduced susceptibility to penicillins, which justifies the fact that amoxicillin and ampicillin are still considered the first line of antibiotic defence against this group of bacteria in both animal [101] and human medicine [1,8,9], respectively. Carbapenems are slightly less efficient towards enterococci, and they are considered resistant to cephalosporins [1,3,8,9].
Bacterial D,D-transpeptidases, also known as penicillin binding proteins (PBPs), are responsible for the last step of cross-linkage in peptidoglycan formation, a complex structure responsible for the stability and rigidness of the bacterial cell wall. β-lactams have the capability of acting as substrates for these PBPs, effectively inhibiting them and, consequently, preventing cell wall formation [102].
In enterococci, intrinsic resistance is usually associated with the existence of low-affinity PBPs that do not allow binding of the antibiotic molecules as easily. The presence of these proteins has been associated with low-level resistance to penicillins and moderate to high-level resistance to cephalosporins [3,8]. PBPs are usually divided into high-molecular and low-molecular weight PBPs. From both these groups, high-molecular PBPs are the ones associated with β-lactam resistance in bacteria [103]. These transpeptidases can also be divided into two classes: class A that presents both transpeptidase and transglycosylase activity, and class B that presents only transpeptidase activity [104].
It has been known for a long time that Enterococcus strains have at least five PBPs [105], associated with six putative genes: three of those genes belong to Class A (ponA, pbpF, pbpZ), and three to Class B (pbp5, pbpA, pbpB) [103,104]. The chromosomally encoded pbp5 gene, present in E. faecium strains, has been the gene most frequently associated to penicillin and cephalosporin resistance in Enterococcus [106]. Although this gene has also been identified in E. faecalis (pbp4), it has not been as associated to β-lactam resistance in this species as in E. faecium [104]. In fact, E. faecalis usually presents higher rates of susceptibility to β-lactams in comparative studies [107,108,109]. It has also been proven that the pbp5 gene is a transferable element between E. faecium and is sometimes associated with the transference of vacomycin-resistance determinants [110,111,112].
A higher resistance to these antibiotics is usually associated with a higher expression of PBP5 proteins [113,114] or to mutations in the amino acid gene sequence that lead to alterations of this protein’s molecular structure [114,115,116]. When evaluating the relation between pbp5 sequences of E. faecium and penicillin’s minimal inhibition concentration (MIC) presented by these isolates, a study by Galloway-Peña et al. [115], corroborated later by another study by Pietta et al. [116], concluded that isolates that presented a MIC ≤ 2 µg/mL and isolates that presented a MIC ≥ 16 µg/mL had a 5% difference in the pbp5 sequence, which remounted to the presence of two distinct allelic forms: a more susceptible (PBP5-S) and a more resistant (PBP5-R) form. These authors also proposed the existence of a hybrid PBP5 (PBP-S/R) for the isolates that present a MIC ≈ 4 µg/mL, whose genomic sequence fell between the other two genes. These differences were also associated to different enterococci clades previously mentioned, with PBP5-R being more associated to Clade A1 and PBP5-S to Clade B [113,115,116]. However, both studies concluded that the determination of which amino acid changes were responsible for this decrease in susceptibility to β-lactams was difficult. The divergence seen in results described in other studies also reinforces this conclusion, although some changes such as an additional serine after amino acid position 466 (Ser-466) or amino acid substitutions (for instance the substitution of a methionine, Met-485, for an alanine or threonine (Met-485 → Ala/Thr), the substitution of a glutamic acid, Glu-629, for a valine, and also the substitution of an aspartic acid, Asp-496, for a lysine) have been more frequently associated to a high-level of penicillin resistance than others [114,115,117,118,119]. Although it has been proven that this PBP is responsible for some degree of resistance to penicillins, high-level resistance cannot be solely justified by the increased expression or mutations of the pbp5 gene, which means that other factors should also play some kind of role in this adaptation [106,114].
On rare occasions, decreased susceptibility to β-lactams has been also associated with the production of β-lactamases, with Murray reporting the first E. faecalis β-lactamase producer in 1983 [120] and Coudron et al. [121] reporting the first E. faecium β-lactamase producer in 1992. Since then, reports on β-lactamase producing enterococci have been scarce and more frequently associated with E. faecalis [1,122]. β-lactamases act through β-lactam ring hydrolysis, originating a molecule incapable of binding to the PBP [102]. Even though Enterococcus and Staphylococcus present a very similar sequenced operon, which indicates the possibility that these genes were transferred from one genus to the other [122], the blaZ gene in enterococci has a low-level constitutive expression, which means that it occurs independently of the presence of β-lactams [104]. This means that an in vitro susceptibility may not be equivalent to an in vivo one, since infections by these bacteria equivalate to higher concentrations of β-lactamases than the ones seen in vitro [3,104]. However, considering that the isolation frequency of these bacteria is reportedly low and also that these β-lactamases seem to be susceptible to β-lactamases inhibitors, such as sulbactam [104,122], resistance to β-lactams through this mechanism is not as concerning as through PBPs.

4.2. Aminoglycosides

Aminoglycosides act through disruption of mRNA decoding by binding to the 16S rRNA of the 30S ribosomal subunit [9]. This means that in order to exert their action, they must first pass through the bacterial cell wall.
Enterococci are known to present an intrinsic resistance to aminoglycosides due to two distinct factors: poor uptake of this antibiotic group through the cell wall, and through modification of the antibiotic molecule, reducing its affinity to their target [3,8,9,104].
In aminoglycosides, modification of the antibiotic molecule occurs due to a group of enzymes that can be divided into three different categories: the acetyl-coenzyme A-dependent amino-glycoside acetyltransferases (AACs), that act through acetylation of the amino groups; the ATP-dependent nucleotidyltransferases (ANTs), that act through adenylation of the hydroxyl group; and the ATP/GTP-dependent phosphotransferases (APHs), that act through phosphorylation of the hydroxyl group [9,123]. These molecules are also divided into sub-groups according to which position is altered (3′, 6′, 2″,3″, 4″, etc.), and according to the resistance profile presented (I, II, III, etc.). Letters (a, b, c, d, e, etc.) were added at the end in order to distinguish each specific protein [1,123,124].
In enterococci, there have been three similar transferases described as being chromosomally encoded and that confer low- to moderate-level resistance to aminoglycosides: AAC(6′)-Ii in E. faecium [125,126,127]; AAC(6′)-Id in E. durans; and AAC(6′)-Ih in E. hirae [127], conferring resistance to kanamycin and tobramycin [126]. Some aminoglycoside-modifying enzymes have also been associated to acquired resistance, by the obtention of genes encoded in transposons and conjugative plasmids. All these enzymes are presented in Table 1, along with the type of resistance they confer.
From all the presented enzymes, AAC-6′-Ie-APH-2 seems to be the most frequently associated with high-level resistance to gentamicin, and ANT(6′)-Ia to high-level resistance to streptomycin [138,139,140,141,142,143,144]. It is important to note, however, that a discordance between phenotype and genotype can occur when it comes to these enzymes, which means that not all bacteria that present a resistance gene for aminoglycosides presents a resistant phenotype [145]. This means that phenotype determination is always important in order to determine resistance to these antibiotics.
Enterococci with the ability to produce these enzymes, conferring high-level resistance to both gentamicin and streptomycin which are the go-to aminoglycosides to control enterococcal infection, are also usually resistant to the synergism between these antibiotics and cell wall active agents such as β-lactams [1,8,9].
Another noteworthy reference is the fact that there is a possibility that, in polymicrobial infections, modifying enzyme-producing enterococci could possibly shield other bacteria from aminoglycoside action, increasing the MIC values necessary to inhibit bacterial multiplication [146].
Intrinsic resistance to tobramycin and kanamycin, in both E. faecium and E. faecalis, is associated to a rRNA methyltransferase, EfmM, that acts through alteration of the ribosomal target-site [147].
High-level resistance to streptomycin has also been associated to punctual ribosome mutations [3,8,148].
Additionally, EfrAB, an ABC-transporter efflux pump, also seems to be responsible for the reduction of gentamicin’s MIC levels in enterococci [149].

4.3. Glycopeptides

It is known that glycopeptides, such as vancomycin and teicoplanin, act by binding to the peptidoglycan pentapeptide precursor, more specifically to the D-alanine-D-alanine (D-Ala-D-Ala) terminus, which blocks the last crosslinking step of peptidoglycan formation, consequently preventing cell wall formation [3,150]. In enterococci, the main mechanism of resistance to these antibiotics is the alteration of the terminus molecule, D-Ala-D-Ala, which corresponds to the glycopeptides target-site, thus reducing the affinity of these antibiotics to these targets [3,9,150].
The operons responsible for this modification are normally divided into two groups, according to the alteration they engender: vanA, vanB, vanD and vanM operons, which lead to the creation of a D-alanine-D-lactate (D-Ala-D-Lac) terminus associated to both vancomycin and teicoplanin moderate- to high-level resistance [150,151,152,153,154]; and the vanC, vanE, vanG, vanL and vanN operons responsible for the creation of a D-alanine-D-serine (D-Ala-D-Ser) terminus and associated only to low-level vancomycin resistance [150,151,153,154,155,156,157,158].
E. faecium is the most common species associated to glycopeptide resistance and vanA is the most described operon, with vanB coming in second; both are present in transposons that are either chromosomally encoded or transferable through plasmids, namely Tn1546 for vanA, and Tn1547 or Tn1549 for vanB [1,3,8,104,150,151].
The vanA operon is composed by seven different genes: vanR, vanS, vanH, vanA, vanX, vanY and vanZ, all working in sequence in order to develop glycopeptide resistance (Figure 2) [1,8,9,104,153].
Regulation of these genes is performed by the VanR-VanS system. In the absence of glycopeptides, the sensor kinase, VanS, acts as an inhibitor to the response regulator, VanR [8,157]. However, when glycopeptides are present in the environment, VanS activates the vanR gene through phosphorylation [1,9]. This gene subsequently activates a promoting region responsible for the transcription of three other genes: vanH, vanA and vanX. It also increases the regulation of the VanR-VanS system [8,159]. Moreover, vanH originates a dehydrogenase that reduces pyruvate to D-Lac that, along with the ligase produced by vanA, promotes the formation of the D-Ala-D-Lac terminus [9,104,153]. The vanX gene is responsible for the formation of a dipeptidase and vanY encodes a carboxypeptidase, both responsible for cleaving the remaining D-Ala-D-Ala forms. While VanX cleaves the free D-Ala-D-Ala terminus before its connection to the peptidoglycan precursors, VanY cleaves this same terminus after the pentapeptides’ formation [9,104,153,160,161]. On another end, vanZ has a still unknown mechanism of action that is somehow related to teicoplanin resistance [6,162].
The vanB operon is very similar to the vanA (Figure 1), with only two major differences: first in the regulatory system, which is encoded by vanRB and vanSB; and second in the absence of the vanZ gene, which is exchanged for the vanW gene, and the reason why enterococci presenting resistance to glycopeptides due to this operon are susceptible to teicoplanin [3,153].
Other operons have been described in enterococci, such as vanC, which is associated with intrinsic resistance in E. gallinarium (vanC1), E. casseliflavus (vanC2) and E. flavescens (vanC3) [3,153,163,164]. Other operons are of scarcer frequency in enterococci [8], and have been described extensively elsewhere [151,153,165].

4.4. Fluoroquinolones

DNA replication is essential in order for cell division to occur. This replication is dependent of the activity of two enzymes: DNA gyrase, responsible for the negative supercoiling of DNA and essential for transcription occurrence; and topoisomerase IV, responsible for the disentanglement of the newly replicated DNA and segregation of the daughter chromosomes [104,166,167]. Both of these enzymes are heterotetramers with two different subunits—GyrA2GyrB2 for DNA gyrase and ParC2ParE2 for topoisomerase IV—and are formed due to the expression of two different groups of genes, gyrA/gyrB and parC/parE, respectively [166,167].
Quinolones act by binding and inhibiting both of these enzymes, which impedes DNA replication and consequently cell division [9,166]. Initial reports associated enterococci resistance to quinolones mainly to mutations in the gyrA gene [168,169,170]. In 1998, Kanematsu et al. [171] also corroborated the fact that these resistances arise from mutations in the gyrA gene, but additionally indicated that they were also associated to mutations in the parC gene. Since then, numerous studies have associated high-level quinolone resistance in E. faecium and E. faecalis to mutations in both of these genes [172,173,174,175,176]. These mutations alter the affinity of these enzymes to quinolones, leading to an impediment in the formation of the quinolone-enzyme-DNA complexes, responsible for the inhibition of cell division [104,166].
Additionally, Kanematsu et al. [171] also pointed out the fact that some isolates presenting low-level resistance to quinolones only present mutations in the parC gene, which could be an indication that points to topoisomerase IV being the primary target of quinolones in enterococci. This seems to be the general case in Gram-positive bacteria, contrary to what happens in Gram-negative bacteria, in which DNA gyrase is considered generally the primary target [8,166,167,171]. This is also corroborated by the fact that Gram-positive DNA gyrase appears to be less susceptible to quinolones when compared to the one of Gram-negative bacteria [166]. However, Oyamada et al. [177] also indicated that the primary target could change according to the quinolone used.
On the other hand, mutations in the gyrB and parE have been very infrequent in these bacteria [178].
Low-level resistance to quinolones in enterococci has also been associated to the activity of efflux pumps [178], mainly EmeA, a MDR efflux pump homolog to Staphylococcus aureus NorA [179].
Finally, in 2007, Arsène and Leclercq [167] also found, in E. faecalis, a protein homologue to QnrA, which they named Qnr E. faecalis. Similarly to QnrA, which is responsible for the protection of DNA gyrase and topoisomerase IV from fluoroquinolone inhibition, Qnr E. faecalis seems to be associated to resistance to ciprofloxacin through DNA gyrase protection [167,180,181,182].

4.5. Tetracyclines

Tetracyclines act through inhibition of protein synthesis by binding to the ribosome and preventing tRNA linking, causing a bacteriostatic effect upon the cell [104].
Resistance to this group of antibiotics has been extensively described in a great number of bacteria and, in enterococci, it has mostly been related either to two genes associated to efflux pumps, tet(K) and tet(L), or to three genes associated to ribosomal protection, tet(M), tet(O) and tet(S) [3]. Of all five genes, tet(M) seems to be the most frequently isolated from enterococci strains from both human and animal origin, with tet(L) coming in close second [142,173,183,184,185,186].
Transference of these genes through the Tn916/Tn1545 transposon family has also been associated to the co-transference of the gene erm(B), associated with macrolide, lincosamide and streptogramin B (MLSB) resistance [3,183,184,187].
Recently, resistance to tetracyclines in various Gram-positive bacteria including enterococci has also been associated to an ATP-binding cassette (ABC)-F ribosomal protection protein, encoded by the poxtA gene. This gene can either be present in the chromosome or be encoded in a transferable plasmid, and is also associated to co-resistance to phenicol and oxazolidinones [188,189,190,191,192].

4.6. Oxazolidinones

Oxazolidinones, especially linezolid, are one of the antibiotics used against vancomycin resistant enterococci (VRE) [193,194,195,196]. Although the percentage of resistance seen in enterococci to this antibiotic is still low (<1%), it has been progressively increasing in the last few years [157,192,197,198,199,200,201,202].
Linezolid acts by binding to the 23s rRNA blocking tRNA docking, and thereby preventing protein translation [104,203].
The most common mechanism of resistance to this antibiotic, especially in E. faecium, consists of point mutations in the 23S rDNA, with the most frequent being the G2576T (E. coli numbering) [198,201,204]. While E. faecium has six genes encoding for this rRNA, E. faecalis has four, and the number of genes that suffer from this mutation has been associated to the degree of resistance presented by these bacteria [8,104,205].
Other mutations, including in ribosomal stabilization proteins L3 (rplC), L4 (rplD) and L22 (rplV), have also been described; however, they seem to be very rare in enterococci [204].
Acquisition of linezolid resistance is also associated to Cfr and Cfr-like methylases (in Enterococcus case cfr(B) and cfr(D)) [192,203,204,206,207,208,209,210,211], ABC-F proteins optrA [197,199,201,212,213] and the previously mentioned PoxtA [192,200,204]. The optrA gene seems to be the most frequent mechanism of resistance to linezolid in E. faecalis [204,212].
Cfr proteins confer resistance through post-transcriptional methylations of the 23S rRNA, which alter susceptibility not only to oxazolidinones, but also to phenicols, lincosamides, pleuromutilins and streptogramin A (PhLOPSA phenotype) [207]. However, the presence of these genes in enterococci does not always equal a resistance phenotype to linezolid [209,211].
OptrA and PoxtA lead to decreased susceptibility to phenicols and oxazolidinones, with PoxtA also conferring resistance to tetracyclines [188,189,197]. While optrA and poxtA are both usually acquired through plasmids, the poxtrA gene can also be chromosome encoded [192].

5. Biocide Tolerance

Biocides have been used for a long time with the intent of reducing the quantity of microorganisms present in different surfaces, and are helpful in the prevention of the growing quantity of multi-resistant organisms, the spread of infections and, consequently, the amount of HAIs occurring in today’s practice. Regulation (EU) no 528/2012 of the European Parliament and the Council of 22 May 2012 defines “biocidal product” as a compound that contains in its composition (or that leads to the formation of) one or more active substances, utilized with the intent of “destroying, deterring or rendering harmless” microorganisms (by other means besides physical or mechanical ones), in order to attenuate or eliminate any detrimental action these agents may have towards host health. These compounds are usually divided into four categories: antiseptics, sterilants, disinfectants and preservatives; however, several compounds can fit into more than one category [214].
Although the term “resistance” is widely used in the “antibiotic world”, the same cannot be said for biocides. In relation to this group of compounds, terms such as “reduced susceptibility” or “tolerance” are more frequently used since while they are associated with increases in the minimal inhibitory/bactericidal concentrations (MICs/MBCs) needed to either inhibit or kill a certain bacterium, they also imply that in-use concentrations are still effective against these microorganisms [215]. This definition, however, becomes a little more unclear when we contemplate the lack of standardized laboratory methods that can be used to determine biocide susceptibility, and MIC/MBC breakpoints that define the line between biocide tolerance and susceptibility [216].
There are a variety of factors that can affect a biocide’s efficiency and lead to bacterial tolerance. These factors can be related to the biocide itself, such as its concentration, pH and formulation; to the treatment conditions in which these compounds are applied, such as the temperature, presence of organic matter and contact time; or to the targeted microorganisms, due to differences in the cell wall, the presence of efflux mechanisms or enzymatic degradation. The presence of organisms organized in the biofilm form are also associated to a higher biocide tolerance [217]. When considering biocide tolerance, concentration is usually considered the most important factor [215,217]; nevertheless, all other factors should be considered in order to achieve an optimal biocidal efficiency and reduce possible decreases in susceptibility.
In enterococci, such as in other bacteria, increases in tolerance to biocides could be associated to protective stress-associated mechanisms triggered when in the presence of sub-lethal concentrations of these compounds, either due to a wrongful application or to biocidal residues left in the environment after usage [16].
Small RNAs (sRNAs), typically composed of noncoding transcripts between 50 and 600 nucleotides, are usually produced under specific environmental conditions [218]. These sRNAs, still being studied in both E. faecalis [219,220] and E. faecium [218], have been thought to act as a regulatory system of a network of genes [219], activated when under antibiotic [218] or biocide [221] stress, possibly leading to an adaptation to both of these antimicrobials and consequent decreases in susceptibility.
Phosphotransferase systems (PTS), responsible for the transportation and phosphorylation of sugars used by bacteria to produce energy, have also been proven to be important in E. faecalis survival to a variety stress conditions [222]. This system has been indicated by Pidot et al. [14] as a possible explanation for an increase in alcohol tolerance in E. faecium.
On the other hand, the two-component system ChtRS, that also reacts to environmental alterations and is coded by the chts and chtr genes that putatively encode a histidine kinase and a response regulator, has been associated to chlorhexidine tolerance in E. faecium [12].
Efflux systems, such as QacA/B and EfrAB, have equally been associated to biocide tolerance in enterococci, especially to chlorhexidine [149,223].
Moreover, it seems valuable to indicate that different enterococci species seem to present different tolerance levels to different biocides, with the most tolerant species variating according to the biocide tested [224]. This means that when studying the efficacy of a biocide against Enterocccus or bacterial tolerance in this genus, different species should be tested.
There has also been a great debate on whether there is a possible association between biocides and antibiotics resistance, with some studies indicating this possibility [14,225], while other studies conclude that a link between the two does not exist [217,223]. What seems to be mostly accepted is that the existence of this association probably depends on the antibiotic and biocide tested and the resistance mechanisms associated. However, the application of sub-lethal concentrations of biocides could possibly co-select antimicrobial-resistant enterococci [226].
This resilience and capability of adaptation to stressful conditions presented by the Enterococcus genus is very much a concern, and it goes beyond antibiotics and biocides, with reports indicating possible decreases of susceptibility even to venoms [227].
Table 2 presents a summary of the main mechanisms of resistance and decreased susceptible to antibiotics and biocides, presented by the Enterococcus genus.

6. COVID-19 and Enterococcus

One of the most menacing HAIs of the last few years has been the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) infection. This infection, like most viral infections, is known to lead to patient immunosuppression, and consequently give-way to secondary bacterial infections caused by commensal opportunistic pathogens [228].
However, among all the opportunistic pathogens, enterococci seem to strive in the presence of SARS-CoV-2 infected patients, especially when it comes to bloodstream infections (BSI) [229,230]. This relation is still not completely clear [229], with some studies indicating that these enterococci could be of nosocomial origin [231], while others indicate that they could originate from the individual himself, due to disruption of the intestinal barrier caused by viral multiplication and a consequent increase of enterococcal concentration in the gut [232].
Considering all these factors, it is also important to note that during the COVID-19 pandemic there was an increase in antimicrobial usage in order to fight these secondary infections, which means that a corresponding increase in bacterial resistance to these compounds is also to be expected [229]. This has been reflected in the enterococcal population with reports of increments in highly resistant strains of Enterococcus [233], including VRE [234].
Although the SARS-CoV-2 infection’s consequences and association with enterococcal infection have been, and are still being, thoroughly studied in human medicine, this has not been the case in veterinary medicine. It has been proven that both cats and dogs can become infected [235]; however, to the best of our knowledge, the consequences of these infections have not been studied in relation to the enterococcal population. Nevertheless, we hypothesize that these consequences could be similar to those seen in human medicine and should be carefully studied in order to understand what kind of impact they could possibly have in combating enterococcal infections.
This means that although enterococci have been gaining relevance as an infection-causing pathogen, especially due to all the factors presented in this review, they become even more a point of concern when considering their increased relevance due to the COVID-19 pandemic.

7. Conclusions

Enterococci seem to have developed a variety of mechanisms that make them more apt at survival in the hospital environment. This capacity seems to be a result of the combination of diverse factors.
Although this genus is not known for its variability of virulence factors, all those described in this review seem to contribute to the development of infection and also to the increased resilience in the presence of adverse conditions, especially due to cell surface proteins, such as Esp, that aid in biofilm formation, structures known for their resistance not only to antibiotics but also to biocides.
On the other hand, the intrinsic resistance these bacteria present to a multitude of antibiotics, associated with their genome plasticity and consequent capability of acquiring new genetic elements connected to this same resistance, makes them a concern in terms of antibiotic therapy, especially in terms of vancomycin-resistant enterococci.
Finally, although representing a field that requires further studying, their multiplicity of stress survival-related mechanisms could make them less susceptible to biocides and consequently able to last in the environment.
All of the characteristics described in this article associated with a possible potentiation by COVID-19 of this genus, which make them a number one priority in terms of vigilance and infection control.

Author Contributions

Conceptualization, C.G., S.G. and M.O.; methodology, C.G., S.G. and M.O.; validation, L.T., S.G. and M.O.; formal analysis, C.G.; investigation, C.G.; resources, L.T., S.G. and M.O.; data curation, C.G.; writing—original draft preparation, C.G.; writing—review and editing, L.T., S.G. and M.O.; supervision, S.G. and M.O.; project administration, S.G. and M.O.; funding acquisition, L.T., S.G. and M.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by CIISA—Centro de Investigação Interdisciplinar em Sanidade Animal, Faculdade de Medicina Veterinária, Universidade de Lisboa, Lisboa, Portugal, Project UIDB/00276/2020; and by Laboratório Associado para Ciência Animal e Veterinária (AL4AnimalS), LA/P/0059/2020–AL4AnimalS (funded by FCT).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hollenbeck, B.L.; Rice, L.B. Intrinsic and acquired resistance mechanisms in Enterococcus. Virulence 2012, 3, 421–433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Lebreton, F.; Willems, R.J.L.; Gilmore, M.S. Enterococcus diversity, origins in nature, and gut colonization. In Enterococci: From Commensals to Leading Causes of Drug Resistant Infection; Gilmore, M.S., Clewell, D.B., Yasuyoshi, I., Shankar, N., Eds.; Eye and Ear Infirmary: Boston, MA, USA, 2014; pp. 5–64. [Google Scholar]
  3. Torres, C.; Alonso, C.A.; Ruiz-Ripa, L.; León-Sampedro, R.; del Campo, R.; Coque, T.M. Antimicrobial resistance in Enterococcus spp. of animal origin. Microbiol. Spectr. 2018, 6, 185–227. [Google Scholar] [CrossRef] [PubMed]
  4. Brinkwirth, S.; Ayobami, O.; Eckmanns, T.; Markwart, R. Hospital-acquired infections caused by enterococci: A systematic review and meta-analysis, WHO European Region, 1 January 2010 to 4 February 2020. Euro Surveill. 2021, 26, 1–16. [Google Scholar] [CrossRef] [PubMed]
  5. Britt, N.S.; Potter, E.M. Clinical epidemiology of vancomycin-resistant Enterococcus gallinarum and Enterococcus casseliflavus bloodstream infections. J. Glob. Antimicrob. Resist. 2017, 5, 57–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Toc, D.A.; Pandrea, S.L.; Botan, A.; Mihaila, R.M.; Costache, C.A.; Colosi, I.A.; Junie, L.M. Enterococcus raffinosus, Enterococcus durans and Enterococcus avium Isolated from a Tertiary Care Hospital in Romania—Retrospective Study and Brief Review. Biology 2022, 11, 598. [Google Scholar] [CrossRef] [PubMed]
  7. Monticelli, J.; Knwzevich, A.; Luzzati, R.; Di Bella, S. Clinical management of non-faecium non-faecalis vancomycin-resistant enterococci infection. Focus on Enterococcus gallinarium and Enterococcus casseliflavus/flavescen. J. Infect. Chemother. 2018, 24, 237–246. [Google Scholar] [CrossRef] [Green Version]
  8. García-Solache, M.; Rice, L.B. The Enterococcus: A model of adaptability to its environment. Clin. Microbiol. Rev. 2019, 32, e00058-18. [Google Scholar] [CrossRef] [Green Version]
  9. Kristich, C.J.; Rice, L.B.; Arias, C.A. Enterococcal infection—Treatment and antibiotic resistance. In Enterococci: From Commensals to Leading Causes of Drug Resistant Infection; Gilmore, M.S., Clewell, D.B., Yasuyoshi, I., Shankar, N., Eds.; Eye and Ear Infirmary: Boston, MA, USA, 2014; pp. 123–184. [Google Scholar]
  10. WHO Publishes List of Bacteria for Which New Antibiotics Are Urgently Needed. Available online: https://www.who.int/news/item/27-02-2017-who-publishes-list-of-bacteria-for-which-new-antibiotics-are-urgently-needed (accessed on 24 January 2022).
  11. Heller, L.C.; Edelblute, C.M. Long-term metabolic persistence of gram-positive bacteria on health care-relevant plastic. Am. J. Infect. Control 2018, 46, 50–53. [Google Scholar] [CrossRef]
  12. Grund, A.; Rautenschlein, S.; Jung, A. Tenacity of Enterococcus cecorum at different environmental conditions. J. Appl. Microbiol. 2020, 130, 1494–1507. [Google Scholar] [CrossRef]
  13. Prieto, A.M.G.; Wijngaarden, J.; Braat, J.C.; Rogers, M.R.C.; Majoor, E.; Brouwer, E.C.; Zhang, X.; Bayjanov, J.R.; Bonten, M.J.M.; Willems, R.J.L.; et al. The two-component system ChtRS contributes to chlorhexidine tolerance in Enterococcus faecium. Antimicrob. Agents Chemother. 2017, 61, e02122-16. [Google Scholar] [CrossRef] [Green Version]
  14. Pidot, S.; Gao, W.; Buultjens, A.; Monk, I.; Guerillot, R.; Carter, G.; Lee, J.; Lam, M.; Grayson, M.L.; Ballard, S.; et al. Increasing tolerance of hospital Enterococcus faecium to hand-wash alcohols. Sci. Transl. Med. 2018, 10, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Sobhanipoor, M.H.; Ahmadrajabi, R.; Nave, H.H.; Saffari, F. Reduced susceptibility to biocides among enterococci from clinical and non-clinical sources. Infect. Chemother. 2021, 53, 696–704. [Google Scholar] [CrossRef] [PubMed]
  16. Geraldes, C.; Verdial, C.; Cunha, E.; Almeida, V.; Tavares, L.; Oliveira, M.; Gil, S. Evaluation of a biocide used in the biological isolation and containment unit of a veterinary teaching hospital. Antibiotics 2021, 10, 639. [Google Scholar] [CrossRef]
  17. Kalchayanand, N.; Koohmaraie, M.; Wheeler, T.L. Effect of exposure time and organic matter on efficacy of antimicrobial compounds against shiga toxin–producing Escherichia coli and Salmonella. J. Food Prot. 2016, 79, 561–568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Weigel, L.M.; Clewell, D.B.; Gill, S.R.; Clark, N.C.; Mcdougal, L.K.; Flannagan, S.E.; Kolonay, J.F.; Shetty, J.; Killgore, G.E.; Tenover, F.C. Genetic analysis of a high-level vancomycin-resistant isolate of Staphylococcus aureus. Science 2003, 302, 1569–1571. [Google Scholar] [CrossRef] [PubMed]
  19. de Niederhäusern, S.; Sabia, C.; Messi, P.; Guerrieri, E.; Manicardi, G.; Bondi, M. Glycopeptide-resistance transferability from vancomycin-resistant enterococci of human and animal source to Listeria spp. Lett. Appl. Microbiol. 2004, 39, 483–489. [Google Scholar] [CrossRef] [PubMed]
  20. Zhu, W.; Clark, N.C.; McDougal, L.K.; Hageman, J.; McDonald, L.C.; Patel, J.B. Vancomycin-resistant Staphylococcus aureus isolates associated with Inc18-like vanA plasmids in Michigan. Antimicrob. Agents Chemother. 2008, 52, 452–457. [Google Scholar] [CrossRef] [Green Version]
  21. Limbago, B.M.; Kallen, A.J.; Zhu, W.; Eggers, P.; McDougal, L.K.; Albrecht, V.S. Report of the 13th vancomycin-resistant Staphylococcus aureus isolate from the United States. J. Clin. Microbiol. 2014, 52, 998–1002. [Google Scholar] [CrossRef] [Green Version]
  22. van Schaik, W.; Top, J.; Riley, D.R.; Boekhorst, J.; Vrijenhoek, J.E.P.; Schapendonk, C.M.E.; Hendrickx, A.P.A.; Nijman, I.J.; Bonten, M.J.M.; Tettelin, H.; et al. Pyrosequencing-based comparative genome analysis of the nosocomial pathogen Enterococcus faecium and identification of a large transferable pathogenicity island. BMC Genom. 2010, 11, 239. [Google Scholar] [CrossRef] [Green Version]
  23. Paulsen, I.; Banerjei, L.; Myers, G.; Nelson, K.; Seshadri, R.; Read, T.; Fouts, D.; Eisen, J.; Gill, S.; Heidelberg, J.; et al. Role of mobile DNA in the evolution of vancomycin-resistant Enterococcus faecalis. Science 2003, 299, 2071–2074. [Google Scholar] [CrossRef] [Green Version]
  24. Ruiz-Garbajosa, P.; Bonten, M.J.M.; Robinson, D.A.; Top, J.; Nallapareddy, S.R.; Torres, C.; Coque, T.M.; Canto, R.; Baquero, F.; Murray, B.E.; et al. Multilocus sequence typing scheme for Enterococcus faecalis reveals hospital-adapted genetic complexes in a background of high rates of recombination. J. Clin. Microbiol. 2006, 44, 2220–2228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Leavis, H.L.; Willems, R.J.L.; van Wamel, W.J.B.; Schuren, F.H.; Caspers, M.P.M.; Bonten, M.J.M. Insertion sequence-driven diversification creates a globally dispersed emerging multiresistant subspecies of E. faecium. PLoS Pathog. 2007, 3, 75–96. [Google Scholar] [CrossRef] [PubMed]
  26. Galloway-Peña, J.; Roh, J.H.; Latorre, M.; Qin, X.; Murray, B.E. Genomic and SNP analyses demonstrate a distant separation of the hospital and community-associated clades of Enterococcus faecium. PLoS ONE 2012, 7, e30187. [Google Scholar] [CrossRef] [PubMed]
  27. Palmer, K.L.; Godfrey, P.; Griggs, A.; Kos, V.N.; Zucker, J.; Desjardins, C.; Cerqueira, G.; Gevers, D.; Walker, S.; Wortman, J.; et 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, e00318-11. [Google Scholar] [CrossRef] [Green Version]
  28. Lebreton, F.; van Schaik, W.; McGuire, A.M.; Godfrey, P.; Griggs, A.; Mazumdar, V.; Corander, J.; Cheng, L.; Saif, S.; Young, S.; et al. Emergence of epidemic multidrug-resistant Enterococcus faecium from animal and commensal strains. MBio 2013, 4, e00534-13. [Google Scholar] [CrossRef] [Green Version]
  29. Raven, K.E.; Reuter, S.; Reynolds, R.; Brodrick, H.J.; Russell, J.E.; Török, M.E.; Parkhill, J.; Peacock, S.J. A decade of genomic history for healthcare-associated Enterococcus faecium in the United Kingdom and Ireland. Genome Res. 2016, 26, 1388–1396. [Google Scholar] [CrossRef] [Green Version]
  30. Zhong, Z.; Kwok, L.Y.; Hou, Q.; Sun, Y.; Li, W.; Zhang, H.; Sun, Z. Comparative genomic analysis revealed great plasticity and environmental adaptation of the genomes of Enterococcus faecium. BMC Genom. 2019, 20, 602. [Google Scholar] [CrossRef]
  31. Solheim, M.; Aakra, Å.; Snipen, L.G.; Brede, D.A.; Nes, I.F. Comparative genomics of Enterococcus faecalis from healthy Norwegian infants. BMC Genom. 2009, 10, 194. [Google Scholar] [CrossRef] [Green Version]
  32. Gao, W.; Howden, B.P.; Stinear, T.P. Evolution of virulence in Enterococcus faecium, a hospital-adapted opportunistic pathogen. Curr. Opin. Microbiol. 2018, 41, 76–82. [Google Scholar] [CrossRef]
  33. Kawalec, M.; Pietras, Z.; Daniłowicz, E.; Jakubczak, A.; Gniadkowski, M.; Hryniewicz, W.; Willems, R.J.L. Clonal structure of Enterococcus faecalis isolated from Polish hospitals: Characterization of epidemic clones. J. Clin. Microbiol. 2007, 45, 147–153. [Google Scholar] [CrossRef] [Green Version]
  34. Neumann, B.; Prior, K.; Bender, J.K.; Harmsen, D.; Klare, I.; Fuchs, S.; Bethe, A.; Zühlke, D.; Göhler, A.; Schwarz, S.; et al. A core genome multilocus sequence typing scheme for Enterococcus faecalis. J. Clin. Microbiol. 2019, 57, e01686-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Leavis, H.L.; Bonten, M.J.; Willems, R.J. Identification of high-risk enterococcal clonal complexes: Global dispersion and antibiotic resistance. Curr. Opin. Microbiol. 2006, 9, 454–460. [Google Scholar] [CrossRef] [PubMed]
  36. Segarra, R.A.; Booth, M.C.; Morales, D.A.; Huycke, M.M.; Gilmore, M.S. Molecular characterization of the Enterococcus faecalis hemolysin/bacteriocin determinant. Infect. Immun. 1991, 59, 1239–1246. [Google Scholar] [CrossRef] [Green Version]
  37. Brock, T.D.; Davie, J.M. Probable identity of a group D hemolysin with a bacteriocine. J. Bacteriol. 1963, 86, 708–712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Ike, Y.; Hashimoto, H.; Clewell, D.B. Hemolysin of Streptococcus faecalis subspecies zymogenes contributes to virulence in mice. Infect. Immun. 1984, 45, 528–530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Gilmore, M.S.; Segarra, R.A.; Booth, M.C. An HlyB-type function is required for expression of the Enterococcus faecalis hemolysin/bacteriocin. Infect. Immun. 1990, 58, 3914–3923. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Miyazaki, S.; Ohno, A.; Kobayashi, I.; Uji, T.; Yamaguchi, K.; Goto, S. Cytotoxic effect of hemolytic culture supernatant from Enterococcus faecalis on mouse polymorphonu-neutrophils and macrophages. Microbiol. Immunol. 1993, 37, 265–270. [Google Scholar] [CrossRef]
  41. Jett, B.D.; Jensen, H.G.; Nordquist, R.E.; Gilmore, M.S. Contribution of the pAD1-encoded cytolysin to the severity of experimental Enterococcus faecalis endophthalmitis. Infect. Immun. 1992, 60, 2445–2452. [Google Scholar] [CrossRef] [Green Version]
  42. Chow, J.W.; Thal, L.A.; Perri, M.B.; Vazquez, J.A.; Donabedian, S.M.; Clewell, D.B.; Zervos, M.J. Plasmid-associated hemolysin and aggregation substance production contribute to virulence in experimental enterococcal endocarditis. Antimicrob. Agents Chemother. 1993, 37, 2474–2477. [Google Scholar] [CrossRef] [Green Version]
  43. Todokoro, D.; Suzuki, T.; Kobayakawa, S.; Tomita, H.; Ohashi, Y.; Akiyama, H. Postoperative Enterococcus faecalis endophthalmitis: Virulence factors leading to poor visual outcome. Jpn. J. Ophthalmol. 2017, 61, 408–414. [Google Scholar] [CrossRef]
  44. Chilambi, G.S.; Nordstrom, H.R.; Evans, D.R.; Kowalski, R.P.; Dhaliwal, D.K.; Jhanji, V.; Shanks, R.M.Q.; van Tyne, D. Genomic and phenotypic diversity of Enterococcus faecalis isolated from endophthalmitis. PLoS ONE 2021, 16, e0250084. [Google Scholar] [CrossRef]
  45. Muller, C.; Breton, Y.; Morin, T.; Benachour, A.; Auffray, Y.; Rincé, A.; Auffray, Y.; Rincé, A. The response regulator CroR modulates expression of the secreted stress induced SalB protein in Enterococcus faecalis. J. Bacteriol. 2006, 188, 2636–2645. [Google Scholar] [CrossRef] [Green Version]
  46. Teng, F.; Kawalec, M.; Weinstock, G.M.; Hryniewicz, W.; Murray, B.E. An Enterococcus faecium secreted antigen, SagA, exhibits broad-spectrum binding to extracellular matrix proteins and appears essential for E. faecium growth. Infect. Immun. 2003, 71, 5033–5041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Pedicord, V.A.; Lockhart, A.A.K.; Rangan, K.J.; Craig, J.W.; Loschko, J.; Rogoz, A.; Hang, H.; Mucida, D. Exploiting a host-commensal interaction to promote intestinal barrier function and enteric pathogen tolerance. Sci. Immunol. 2016, 1, eaai7732. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Rangan, K.J.; Pedicord, V.A.; Wang, Y.-C.; Kim, B.; Lu, Y.; Shaham, S.; Mucida, D.; Hang, H. A secreted bacterial peptidoglycan hydrolase enhances tolerance to enteric pathogens. Science 2016, 353, 1434–1437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Kim, B.; Wang, Y.C.; Hespen, C.W.; Espinosa, J.; Salje, J.; Rangan, K.J.; Oren, D.A.; Kang, J.Y.; Pedicord, V.A.; Hang, H.C. Enterococcus faecium secreted antigen A generates muropeptides to enhance host immunity and limit bacterial pathogenesis. Elife 2019, 8, e45343. [Google Scholar] [CrossRef]
  50. Paganelli, F.L.; de Been, M.; Braat, J.C.; Hoogenboezem, T.; Vink, C.; Bayjanov, J.; Rogers, M.R.C.; Huebner, J. Distinct SagA from hospital-associated clade A1 Enterococcus faecium strains contributes to biofilm formation. Appl. Environ. Microbiol. 2015, 81, 6873–6882. [Google Scholar] [CrossRef] [Green Version]
  51. Park, S.Y.; Kim, K.M.; Lee, J.H.; Seo, S.J.; Lee, I.H. Extracellular gelatinase of Enterococcus faecalis destroys a defense system in insect hemolymph and human serum. Infect. Immun. 2007, 75, 1861–1869. [Google Scholar] [CrossRef] [Green Version]
  52. Thurlow, L.R.; Thomas, V.C.; Narayanan, S.; Olson, S.; Fleming, S.D.; Hancock, L.E. Gelatinase contributes to the pathogenesis of endocarditis caused by Enterococcus faecalis. Infect. Immun. 2010, 78, 4936–4943. [Google Scholar] [CrossRef] [Green Version]
  53. Qin, X.; Singh, K.; Weinstock, G.M.; Murray, B.E. Effects of Enterococcus faecalis fsr genes on production of gelatinase and a serine protease and virulence. Infect. Immun. 2000, 68, 2579–2586. [Google Scholar] [CrossRef] [Green Version]
  54. Qin, X.; Singh, K.; Weinstock, G.M.; Murray, B.E. Characterization of fsr, a regulator controlling expression of gelatinase and serine protease in Enterococcus faecalis OG1RF. J. Bacteriol. 2001, 183, 3372–3382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Bourgogne, A.; Hilsenbeck, S.G.; Dunny, G.M.; Murray, B.E. Comparison of OG1RF and an isogenic fsrB deletion mutant by transcriptional analysis: The Fsr system of Enterococcus faecalis is more than the activator of gelatinase and serine protease. J. Bacteriol. 2006, 188, 2875–2884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Pinkston, K.L.; Gao, P.; Diaz-Garcia, D.; Sillanpää, J.; Nallapareddy, S.R.; Murray, B.E.; Harvey, B.R. The Fsr quorum-sensing system of Enterococcus faecalis modulates surface display of the collagen-binding MSCRAMM Ace through regulation of gelE. J. Bacteriol. 2011, 193, 4317–4325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Rice, L.B.; Carias, L.; Rudin, S.; Vael, C.; Goossens, H.; Konstabel, C.; Klare, I.; Nallapareddy, S.R.; Huang, W.; Murray, B.E. A potential virulence gene, hylEfm, predominates in Enterococcus faecium of clinical origin. J. Infect. Dis. 2003, 187, 508–512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Top, J.; Willems, R.; Bonten, M. Emergence of CC17 Enterococcus faecium: From commensal to hospital-adapted pathogen. FEMS Immunol. Med. Microbiol. 2008, 52, 297–308. [Google Scholar] [CrossRef] [Green Version]
  59. Arias, C.A.; Panesso, D.; Singh, K.; Rice, L.B.; Murray, B.E. Co-transfer of antibiotic resistance genes and a hylEfm-containing virulence plasmid in Enterococcus faecium. Antimicrob. Agents Chemother. 2009, 53, 4240–4246. [Google Scholar] [CrossRef] [Green Version]
  60. Fabretti, F.; Theilacker, C.; Baldassarri, L.; Kaczynski, Z.; Kropec, A.; Holst, O.; Huebner, J.; Infettive, M.; Istituto, I. Alanine esters of enterococcal lipoteichoic acid play a role in biofilm formation and resistance to antimicrobial peptides. Infect. Immun. 2006, 74, 4164–4171. [Google Scholar] [CrossRef] [Green Version]
  61. Geiss-Liebisch, S.; Rooijakkers, S.H.M.; Beczala, A.; Sanchez-Carballo, P.; Kruszynska, K.; Repp, C.; Sakinc, T.; Vinogradov, E.; Holst, O.; Huebner, J.; et al. Secondary cell wall polymers of Enterococcus faecalis are critical for resistance to complement activation via mannose-binding lectin. J. Biol. Chem. 2012, 287, 37769–37777. [Google Scholar] [CrossRef] [Green Version]
  62. Hendrickx, A.P.A.; Willems, R.J.L.; Bonten, M.J.M.; van Schaik, W. LPxTG surface proteins of enterococci. Trends Microbiol. 2009, 17, 423–430. [Google Scholar] [CrossRef]
  63. Isenmann, R.; Schwarz, M.; Rozdzinski, E.; Marre, R.; Beger, H.G. Aggregation substance promotes colonic mucosal invasion of Enterococcus faecalis in an ex vivo model. J. Surg. Res. 2000, 89, 132–138. [Google Scholar] [CrossRef]
  64. Sartingen, S.; Rozdzinski, E.; Muscholl-Silberhorn, A.; Marre, R. Aggregation substance increases adherence and internalization, but not translocation, of Enterococcus faecalis through different intestinal epithelial cells in vitro. Infect. Immun. 2000, 68, 6044–6047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Wells, C.L.; Moore, E.A.; Hoag, J.A.; Hirt, H.; Dunny, G.M.; Erlandsen, S.L. Inducible expression of Enterococcus faecalis aggregation substance surface protein facilitates bacterial internalization by cultured enterocytes. Infect. Immun. 2000, 68, 7190–7194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Kreft, B.; Marre, R.; Schramm, U.; Wirth, R. Aggregation substance of Enterococcus faecalis mediates adhesion to cultured renal tubular cells. Infect. Immun. 1992, 60, 25–30. [Google Scholar] [CrossRef] [Green Version]
  67. Rozdzinski, E.; Marre, R.; Susa, M.; Wirth, R.; Muscholl-Silberhorn, A. Aggregation substance-mediated adherence of Enterococcus faecalis to immobilized extracellular matrix proteins. Microb. Pathog. 2001, 30, 211–220. [Google Scholar] [CrossRef] [PubMed]
  68. Vanek, N.N.; Simon, S.I.; Jacques-Palaz, K.; Mariscalco, M.M.; Dunny, G.M.; Rakita, R.M. Enterococcus faecalis aggregation substance promotes opsonin-independent binding to human neutrophils via a complement receptor type 3-mediated mechanism. FEMS Immunol. Med. Microbiol. 1999, 26, 49–60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Süßmuth, S.D.; Muscholl-Silberhorn, A.; Wirth, R.; Susa, M.; Marre, R.; Rozdzinski, E. Aggregation substance promotes adherence, phagocytosis, and intracellular survival of Enterococcus faecalis within human macrophages and suppresses respiratory burst. Infect. Immun. 2000, 68, 4900–4906. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Rakita, R.M.; Vanek, N.N.; Jacques-Palaz, K.; Mee, M.; Mariscalco, M.M.; Dunny, G.M.; Snuggs, M.; van Winkle, W.B.; Simon, S.I. Enterococcus faecalis bearing aggregation substance is resistant to killing by human neutrophils despite phagocytosis and neutrophil activation. Infect. Immun. 1999, 67, 6067–6075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Paoletti, C.; Foglia, G.; Princivalli, M.S.; Magi, G.; Guaglianone, E.; Donelli, G.; Pruzzo, C.; Biavasco, F.; Facinelli, B. Co-transfer of vanA and aggregation substance genes from Enterococcus faecalis isolates in intra- and interspecies matings. J. Antimicrob. Chemother. 2007, 59, 1005–1009. [Google Scholar] [CrossRef]
  72. Schlievert, P.M.; Gahr, P.J.; Assimacopoulos, A.P.; Dinges, M.M.; Stoehr, J.A.; Harmala, J.W.; Hirt, H.; Dunny, G.M. Aggregation and binding substances enhance pathogenicity in rabbit models of Enterococcus faecalis endocarditis. Infect. Immun. 1998, 66, 218–223. [Google Scholar] [CrossRef] [Green Version]
  73. Bhatty, M.; Cruz, M.R.; Frank, K.L.; Laverde-Gomez, J.A.; Andrade, F.; Garsin, D.A.; Dunny, G.M.; Kaplan, H.B.; Christie, P.J. Enterococcus faecalis pCF10-encoded surface proteins PrgA, PrgB (Aggregation Substance), and PrgC contribute to plasmid transfer, biofilm formation, and virulence. Mol. Microbiol. 2015, 95, 660–677. [Google Scholar] [CrossRef] [Green Version]
  74. Rich, R.L.; Kreikemeyer, B.; Owens, R.T.; LaBrenz, S.; Narayana, S.V.L.; Weinstock, G.M.; Murray, B.E.; Höök, M. Ace is a collagen-binding MSCRAMM from Enterococcus faecalis. J. Biol. Chem. 1999, 274, 26939–26945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Nallapareddy, S.R.; Weinstock, G.M.; Murray, B.E. Clinical isolates of Enterococcus faecium exhibit strain-specific collagen binding mediated by Acm, a new member of the MSCRAMM family. Mol. Microbiol. 2003, 47, 1733–1747. [Google Scholar] [CrossRef] [PubMed]
  76. Sillanpää, J.; Nallapareddy, S.R.; Prakash, V.P.; Qin, X.; Hook, M.; Weinstock, G.M.; Murray, B.E. Identification and phenotypic characterization of a second collagen adhesin, Scm, and genome-based identification and analysis of 13 other predicted MSCRAMMs, including four distinct pilus loci, in Enterococcus faecium. Microbiology 2008, 154, 3199–3211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Nallapareddy, S.R.; Qin, X.; Weinstock, G.M.; Höök, M.; Murray, B.E. Enterococcus faecalis adhesin, Ace, mediates attachment to extracellular matrix proteins collagen type IV and laminin as well as collagen type I. Infect. Immun. 2000, 68, 5218–5224. [Google Scholar] [CrossRef] [Green Version]
  78. Kowalski, W.J.; Kasper, E.L.; Hatton, J.F.; Murray, B.E.; Nallapareddy, S.R.; Gillespie, M.J. Enterococcus faecalis adhesin, Ace, mediates attachment to particulate dentin. J. Endod. 2006, 32, 634–637. [Google Scholar] [CrossRef]
  79. Nallapareddy, S.R.; Singh, K.; Murray, B.E. Contribution of the collagen adhesin Acm to pathogenesis of Enterococcus faecium in experimental endocarditis. Infect. Immun. 2008, 76, 4120–4128. [Google Scholar] [CrossRef] [Green Version]
  80. Singh, K.; Nallapareddy, S.R.; Sillanpää, J.; Murray, B.E. Importance of the collagen adhesin Ace in pathogenesis and protection against Enterococcus faecalis experimental endocarditis. PLoS Pathog. 2010, 6, e1000716. [Google Scholar] [CrossRef]
  81. Singh, K.; Pinkston, K.L.; Gao, P.; Harvey, B.R.; Murray, B.E. Anti-Ace monoclonal antibody reduces Enterococcus faecalis aortic valve infection in a rat infective endocarditis model. Pathog. Dis. 2018, 76, fty084. [Google Scholar] [CrossRef]
  82. Nallapareddy, S.R.; Singh, K.; Okhuysen, P.C.; Murray, B.E. A functional collagen adhesin gene, acm, in clinical isolates of Enterococcus faecium correlates with the recent success of this emerging nosocomial pathogen. Infect. Immun. 2008, 76, 4110–4119. [Google Scholar] [CrossRef] [Green Version]
  83. Tendolkar, P.M.; Baghdayan, A.S.; Shankar, N. Putative surface proteins encoded within a novel transferable locus confer a high-biofilm phenotype to Enterococcus faecalis. J. Bacteriol. 2006, 188, 2063–2072. [Google Scholar] [CrossRef] [Green Version]
  84. Molinos, A.C.; Abriouel, H.; Omar, N.; López, R.L.; Galvez, A. Detection of ebp (endocarditis- and biofilm-associated pilus) genes in enterococcal isolates from clinical and non-clinical origin. Int. J. Food Microbiol. 2008, 126, 123–126. [Google Scholar]
  85. Nallapareddy, S.R.; Singh, K.; Sillanpää, J.; Garsin, D.A.; Höök, M.; Erlandsen, S.L.; Murray, B.E. Endocarditis and biofilm-associated pili of Enterococcus faecalis. J. Clin. Investig. 2006, 116, 2799–2807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Qin, X.; Galloway-Peña, J.R.; Sillanpää, J.; Roh, J.H.; Nallapareddy, S.R.; Chowdhury, S.; Bourgogne, A.; Choudhury, T.; Muzny, D.M.; Buhay, C.J.; et al. Complete genome sequence of Enterococcus faecium strain TX16 and comparative genomic analysis of Enterococcus faecium genomes. BMC Microbiol. 2012, 12, 135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Sillanpää, J.; Nallapareddy, S.R.; Singh, K.; Parakash, V.P.; Fothergill, T.; Ton-That, H.; Murray, B.E. Characterization of the ebpfm pilus-encoding operon of Enterococcus faecium and its role in biofilm formation and virulence in a murine model of urinary tract infection. Virulence 2010, 1, 236–246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Montealegre, M.C.; Singh, K.; Somarajan, S.R.; Yadav, P.; Chang, C.; Spencer, R.; Sillanpää, J.; Ton-That, H.; Murray, B.E. Role of the Emp pilus subunits of Enterococcus faecium in biofilm formation, adherence to host extracellular matrix components, and experimental infection. Infect. Immun. 2016, 84, 1491–1500. [Google Scholar] [CrossRef] [Green Version]
  89. Shankar, V.; Baghdayan, A.S.; Huycke, M.M.; Lindahl, G.; Gilmore, M.S. Infection-derived Enterococcus faecalis strains are enriched in esp, a gene encoding a novel surface protein. Infect. Immun. 1999, 67, 193–200. [Google Scholar] [CrossRef] [Green Version]
  90. Eaton, T.J.; Gasson, M.J. A variant enterococcal surface protein Espfm in Enterococcus faecium; distribution among food, commensal, medical, and environmental isolates. FEMS Microbiol. Lett. 2002, 216, 269–275. [Google Scholar] [CrossRef] [Green Version]
  91. Toledo-Arana, A.; Valle, J.; Solano, C.; Arrizubieta, M.J.; Cucarella, C.; Lamata, M.; Amorena, B.; Leiva, J.; Penadés, J.R.; Lasa, I. The enterococcal surface protein, Esp, is involved in Enterococcus faecalis biofilm formation. Appl. Environ. Microbiol. 2001, 67, 4538–4545. [Google Scholar] [CrossRef] [Green Version]
  92. Tendolkar, P.M.; Baghdayan, A.S.; Gilmore, M.S.; Shankar, N. Enterococcal surface protein, Esp, enhances biofilm formation by Enterococcus faecalis. Infect. Immun. 2004, 72, 6032–6039. [Google Scholar] [CrossRef] [Green Version]
  93. Heikens, E.; Bonten, M.J.M.; Willems, R.J.L. Enterococcal surface protein Esp is important for biofilm formation of Enterococcus faecium E1162. J. Bacteriol. 2007, 189, 8233–8240. [Google Scholar] [CrossRef] [Green Version]
  94. Taglialegna, A.; Matilla-Cuenca, L.; Dorado-Morales, P.; Navarro, S.; Ventura, S.; Garnett, J.A.; Lasa, I.; Valle, J. The biofilm-associated surface protein Esp of Enterococcus faecalis forms amyloid-like fibers. Nat. Partn. J. 2020, 6, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Heikens, E.; Singh, K.; Jacques-Palaz, K.D.; van Luit-Asbroek, M.; Oostdijk, E.A.N.; Bonten, M.J.M.; Murray, B.E.; Willems, R.J.L. Contribution of the enterococcal surface protein Esp to pathogenesis of Enterococcus faecium endocarditis. Microbes Infect. 2011, 13, 1185–1190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Shankar, N.; Lockatell, C.V.; Baghdayan, A.S.; Drachenberg, C.; Gilmore, M.S.; Johnson, D.E. Role of Enterococcus faecalis surface protein Esp in the pathogenesis of ascending urinary tract infection. Infect. Immun. 2001, 69, 4366–4372. [Google Scholar] [CrossRef] [Green Version]
  97. Meredith, K.; Bolhuis, A.; O’Neill, M.A.A. Enterococcal surface protein Esp affects antibiotic sensitivity in Enterococcus faecium. Int. J. Antimicrob. Agents 2009, 34, 392–393. [Google Scholar] [CrossRef] [PubMed]
  98. Weng, P.L.; Ramli, R.; Hamat, R.A. Antibiotic susceptibility patterns, biofilm formation and esp gene among clinical enterococci: Is there any association? Int. J. Environ. Res. Public Health 2019, 16, 3439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Prieto, A.M.G.; Urbanus, R.T.; Zhang, X.; Bierschenk, D.; Koekman, C.A.; van Luit-Asbroek, M.; Ouwerkerk, J.P.; Pape, M.; Paganelli, F.L.; Wobser, D.; et al. The N-terminal domain of the thermo-regulated surface protein PrpA of Enterococcus faecium binds to fibrinogen, fibronectin and platelets. Nat. Publ. Group 2015, 5, 18255. [Google Scholar]
  100. Hendrickx, A.P.A.; van Luit-Asbroek, M.; Schapendonk, C.M.E.; van Wamel, W.J.B.; Braat, J.C.; Wijnands, L.M.; Bonten, M.J.M.; Willems, R.J.L. SgrA, a nidogen-binding LPXTG surface adhesin implicated in biofilm formation, and EcbA, a collagen binding MSCRAMM, are two novel adhesins of hospital-acquired Enterococcus faecium. Infect. Immun. 2009, 77, 5097–5106. [Google Scholar] [CrossRef] [Green Version]
  101. Weese, J.S.; Blondeau, J.; Boothe, D.; Guardabassi, L.G.; Gumley, N.; Papich, M.; Jessen, L.R.; Lappin, M.; Rankin, S.; Westropp, J.L.; et al. International Society for Companion Animal Infectious Diseases (ISCAID) guidelines for the diagnosis and management of bacterial urinary tract infections in dogs and cats. Vet. J. 2019, 247, 8–25. [Google Scholar] [CrossRef]
  102. Lima, L.M.; da Silva, B.N.M.; Barbosa, G.; Barreiro, E.J. β-Lactam antibiotics: An overview from a medicinal chemistry perspective. Eur. J. Med. Chem. 2020, 208, 112829. [Google Scholar] [CrossRef]
  103. Arbeloa, A.; Segal, H.; Hugonnet, J.E.; Josseaume, N.; Dubost, L.; Brouard, J.P.; Gutmann, L.; Mengin-Lecreulx, D.; Arthur, M. Role of class A penicillin-binding proteins in PBP5-mediated β-lactam resistance in Enterococcus faecalis. J. Bacteriol. 2004, 186, 1221–1228. [Google Scholar] [CrossRef] [Green Version]
  104. Miller, W.R.; Munita, J.M.; Arias, C.A. Mechanisms of antibiotic resistance in enterococci. Expert Rev. Anti-Infect. Ther. 2014, 12, 1221–1236. [Google Scholar] [CrossRef] [PubMed]
  105. Williamson, R.; Gutmann, L.; Horaud, T.; Delbos, F.; Acar, J.F. Use of penicillin-binding proteins for the identification of enterococci. J. Gen. Microbiol. 1986, 132, 1929–1937. [Google Scholar] [CrossRef] [PubMed]
  106. Sifaoui, F.; Arthur, M.; Rice, L.; Gutmann, L. Role of penicillin-binding protein 5 in expression of ampicillin resistance and peptidoglycan structure in Enterococcus faecium. Antimicrob. Agents Chemother. 2001, 45, 2594–2597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Ossiprandi, M.C.; Bottarelli, E.; Cattabiani, F.; Bianchi, E. Susceptibility to vancomycin and other antibiotics of 165 Enterococcus strains isolated from dogs in Italy. Comp. Immunol. Microbiol. Infect. Dis. 2008, 31, 1–9. [Google Scholar] [CrossRef]
  108. Ghosh, A.; Dowd, S.E.; Zurek, L. Dogs leaving the ICU carry a very large multi-drug resistant enterococcal population with capacity for biofilm formation and horizontal gene transfer. PLoS ONE 2011, 6, e22451. [Google Scholar] [CrossRef] [Green Version]
  109. Bang, K.; An, J.U.; Kim, W.; Dong, H.J.; Kim, J.; Cho, S. Antibiotic resistance patterns and genetic relatedness of Enterococcus faecalis and Enterococcus faecium isolated from military working dogs in Korea. J. Vet. Sci. 2017, 18, 229–236. [Google Scholar] [CrossRef]
  110. Rice, L.B.; Carias, L.L.; Rudin, S.; Laktičová, V.; Wood, A.; Hutton-Thomas, R. Enterococcus faecium low-affinity pbp5 is a transferable determinant. Antimicrob. Agents Chemother. 2005, 49, 5007–5012. [Google Scholar] [CrossRef] [Green Version]
  111. García-Solache, M.; Lebreton, F.; McLaughlin, R.E.; Whiteaker, J.D.; Gilmore, M.S.; Rice, L.B. Homologous recombination within large chromosomal regions facilitates acquisition of β-lactam and vancomycin resistance in Enterococcus faecium. Antimicrob. Agents Chemother. 2016, 60, 5777–5786. [Google Scholar] [CrossRef] [Green Version]
  112. Novais, C.; Tedim, A.P.; Lanza, V.F.; Freitas, A.R.; Silveira, E.; Escada, R.; Roberts, A.P.; Al-Haroni, M.; Baquero, F.; Peixe, L.; et al. Co-diversification of Enterococcus faecium core genomes and PBP5: Evidences of PBP5 horizontal transfer. Front. Microbiol. 2016, 7, 1581. [Google Scholar] [CrossRef]
  113. Montealegre, M.C.; Roh, J.H.; Rae, M.; Davlieva, M.G.; Singh, K.; Shamoo, Y.; Murray, B.E. Differential penicillin-binding protein 5 (PBP5) levels in the Enterococcus faecium clades with different levels of ampicillin resistance. Antimicrob. Agents Chemother. 2017, 61, e02034-16. [Google Scholar] [CrossRef] [Green Version]
  114. Darehkordi, H.; Saffari, F.; Mollaei, H.R.; Ahmadrajabi, R. Amino acid substitution mutations and mRNA expression levels of the pbp5 gene in clinical Enterococcus faecium isolates conferring high level ampicillin resistance. J. Pathol. Microbiol. Immunol. 2019, 127, 115–122. [Google Scholar]
  115. Galloway-Peña, J.R.; Rice, L.B.; Murray, B.E. Analysis of PBP5 of early U.S. isolates of Enterococcus faecium: Sequence variation alone does not explain increasing ampicillin resistance over time. Antimicrob. Agents Chemother. 2011, 55, 3272–3277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Pietta, E.; Montealegre, M.C.; Roh, J.H.; Cocconcelli, P.S.; Murray, B.E. Enterococcus faecium PBP5-S/R, the missing link between PBP5-S and PBP5-R. Antimicrob. Agents Chemother. 2014, 58, 6978–6981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Mohn, S.C.; Ulvik, A.; Jureen, R.; Willems, R.J.L.; Top, J.; Leavis, H.; Harthug, S.; Langeland, N. Duplex real-time PCR assay for rapid detection of ampicillin-resistant Enterococcus faecium. Antimicrob. Agents Chemother. 2004, 48, 556–560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Poeta, P.; Costa, D.; Igrejas, G.; Sáenz, Y.; Zarazaga, M.; Rodrigues, J.; Torres, C. Polymorphisms of the pbp5 gene and correlation with ampicillin resistance in Enterococcus faecium isolates of animal origin. J. Med. Microbiol. 2007, 56, 236–240. [Google Scholar] [CrossRef]
  119. Belhaj, M.; ben Boubaker, I.B.; Slim, A. Penicillin-binding protein 5 sequence alteration and levels of plp5 mRNA expression in clinical isolates of Enterococcus faecium with different levels of ampicillin resistance. Microb. Drug Resist. 2016, 22, 202–210. [Google Scholar] [CrossRef]
  120. Murray, B.E.; Mederski-Samaroj, B. A new mechanism for in vitro penicillin resistance in Streptococcus faecalis. J. Clin. Investig. 1983, 72, 1168–1171. [Google Scholar] [CrossRef]
  121. Coudron, P.E.; Markowitz, S.M.; Wong, E.S. Isolation of a beta-lactamase-producing, aminoglycoside-resistant strain of Enterococcus faecium. Antimicrob. Agents Chemother. 1992, 36, 1125–1126. [Google Scholar] [CrossRef] [Green Version]
  122. Sarti, M.; Campanile, F.; Sabia, C.; Santagati, M.; Gargiulo, R.; Stefani, A. Polyclonal diffusion of beta-lactamase-producing Enterococcus faecium. J. Clin. Microbiol. 2012, 50, 169–172. [Google Scholar] [CrossRef] [Green Version]
  123. Smith, C.A.; Toth, M.; Stewart, N.K.; Maltz, L.; Vakulenko, S.B. Structural basis for the diversity of the mechanism of nucleotide hydrolysis by the aminoglycoside-2′′-phosphotransferases. Acta Crystallogr. Sect. D Struct. Biol. 2019, 75, 1129–1137. [Google Scholar] [CrossRef]
  124. Shaw, K.J.; Rather, P.N.; Hare, R.S.; Miller, G.H. Molecular genetics of aminoglycoside resistance genes and familial relationships of the aminoglycoside-modifying enzymes. Microbiol. Rev. 1993, 57, 138–163. [Google Scholar] [CrossRef] [PubMed]
  125. Costa, Y.; Galimand, M.; Leclercq, R.; Duval, J.; Courvalin, P. Characterization of the chromosomal aac(6′)-Ii gene specific for Enterococcus faecium. Antimicrob. Agents Chemother. 1993, 37, 1896–1903. [Google Scholar] [CrossRef] [Green Version]
  126. Wright, G.D.; Ladak, P. Overexpression and characterization of the chromosomal aminoglycoside 6′-N-acetyltransferase from Enterococcus faecium. Antimicrob. Agents Chemother. 1997, 41, 956–960. [Google Scholar] [CrossRef] [Green Version]
  127. del Campo, R.; Galán, J.C.; Tenorio, C.; Ruiz-Garbajosa, P.; Zarazaga, M.; Torres, C.; Baquero, F. New aac(6′)-I genes in Enterococcus hirae and Enterococcus durans: Effect on β-lactam/aminoglycoside synergy. J. Antimicrob. Chemother. 2005, 55, 1053–1055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Trieu-Cuot, P.; Courvalin, P. Nucleotide sequence of the Streptococcus faecalis plasmid gene encoding the 3′5″-aminoglycoside phosphotransferase type III. Gene 1983, 23, 331–341. [Google Scholar] [CrossRef]
  129. Kao, S.J.; You, I.; Clewell, D.B.; Donabedian, S.M.; Zervos, M.J.; Petrin, J.; Shaw, K.J.; Chow, J.W. Detection of the high-level aminoglycoside resistance gene aph(2″)-Ib in Enterococcus faecium. Antimicrob. Agents Chemother. 2000, 44, 2876–2879. [Google Scholar] [CrossRef] [Green Version]
  130. Chow, J.W.; Zervos, M.J.; Lerner, S.A.; Thal, L.A.; Donabedian, S.M.; Jaworski, D.D.; Tsai, S.; Shaw, K.J.; Clewell, D.B. A novel gentamicin resistance gene in Enterococcus. Antimicrob. Agents Chemother. 1997, 41, 511–514. [Google Scholar] [CrossRef] [Green Version]
  131. Tsai, S.F.; Zervos, M.J.; Clewell, D.B.; Donabedian, S.M.; Sahm, D.F.; Chow, J.W. A new high-level gentamicin resistance gene, aph(2″)-Id, in Enterococcus spp. Antimicrob. Agents Chemother. 1998, 42, 1229–1232. [Google Scholar] [CrossRef] [Green Version]
  132. Alam, M.M.; Kobayashi, N.; Ishino, M.; Sumi, A.; Kobayashi, K.-I.; Uehara, N.; Watanabe, N. Detection of a novel aph(2″) allele (aph [2″]-Ie) conferring high-level gentamicin resistance and a spectinomycin resistance gene ant(9)-Ia (aad9) in clinical isolates of enterococci. Microb. Drug Resist. 2005, 11, 239–247. [Google Scholar] [CrossRef]
  133. Ounissi, H.; Derlot, E.; Carlier, C.; Courvalin, P. Gene homogeneity for aminoglycoside-modifying enzymes in gram-positive cocci. Antimicrob. Agents Chemother. 1990, 34, 2164–2168. [Google Scholar] [CrossRef] [Green Version]
  134. Carlier, C.; Courvalin, P. Emergence of 4′,4″-aminoglycoside nucleotidyltransferase in enterococci. Antimicrob. Agents Chemother. 1990, 34, 1565–1569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Courvalin, P.; Carlier, C.; Collatz, E. Plasmid-mediated resistance to aminocyclitol antibiotics in group D streptococci. J. Bacteriol. 1980, 143, 541–551. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Mederski-Samoraj, B.D.; Murray, B.E. High-level resistance to gentamicin in clinical isolates of enterococci. J. Infect. Dis. 1983, 147, 751–757. [Google Scholar] [CrossRef]
  137. Ferretti, J.J.; Gilmore, K.S.; Courvalin, P. Nucleotide sequence analysis of the gene specifying the phosphotransferase enzyme in Streptococcus faecalis and identification and cloning of gene regions specifying the two activities. J. Bacteriol. 1986, 167, 631–638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. López, M.; Rezusta, A.; Seral, C.; Aspiroz, C.; Marne, C.; Aldea, M.J.; Ferrer, I.; Revillo, M.J.; Castillo, F.J.; Torres, C. Detection and characterization of a ST6 clone of vanB2-Enterococcus faecalis from three different hospitals in Spain. Eur. J. Clin. Microbiol. Infect. Dis. 2012, 31, 257–260. [Google Scholar] [CrossRef] [PubMed]
  139. ben Said, L.; Klibi, N.; Lozano, C.; Dziri, R.; ben Slama, K.; Boudabous, A.; Torres, C. Diversity of enterococcal species and characterization of high-level aminoglycoside resistant enterococci of samples of wastewater and surface water in Tunisia. Sci. Total Environ. 2015, 530–531, 11–17. [Google Scholar] [CrossRef]
  140. Niu, H.; Yu, H.; Hu, T.; Tian, G.; Zhang, L.; Guo, X.; Hu, H.; Wang, Z. The prevalence of aminoglycoside-modifying enzyme and virulence genes among enterococci with high-level aminoglycoside resistance in Inner Mongolia, China. Braz. J. Microbiol. 2016, 47, 691–696. [Google Scholar] [CrossRef] [Green Version]
  141. Ngbede, E.O.; Raji, M.A.; Kwanashie, C.N.; Kwaga, J.K.P.; Adikwu, A.A.; Maurice, N.A.; Adamu, A.M. Characterization of high-level ampicillin-and aminoglycoside resistant enterococci isolated from non-hospital sources. J. Med. Microbiol. 2017, 66, 1027–1032. [Google Scholar] [CrossRef]
  142. bin Kim, Y.; Seo, K.W.; Son, S.H.; Noh, E.B.; Lee, Y.J. Genetic characterization of high-level aminoglycoside-resistant Enterococcus faecalis and Enterococcus faecium isolated from retail chicken meat. Poult. Sci. 2019, 98, 5981–5988. [Google Scholar]
  143. Harada, S.; Shibue, Y.; Aoki, K.; Ishii, Y.; Tateda, K. Prevalence of high-level aminoglycoside resistance and genes encoding aminoglycoside-modifying enzymes in Enterococcus faecalis and Enterococcus faecium isolated in a University Hospital in Tokyo. Jpn. J. Infect. Dis. 2020, 73, 476–480. [Google Scholar] [CrossRef]
  144. Peyvasti, V.S.; Mobarez, A.M.; Shahcheraghi, F.; Khoramabadi, N.; Rahmati, N.R.; Doust, R.H. High-level aminoglycoside resistance and distribution of aminoglycoside resistance genes among Enterococcus spp. clinical isolates in Tehran, Iran. J. Glob. Antimicrob. Resist. 2020, 20, 318–323. [Google Scholar] [CrossRef] [PubMed]
  145. Adamecz, Z.; Nielsen, K.L.; Kirkby, N.S.; Frimodt-Møller, N. Aminoglycoside resistance genes in Enterococcus faecium: Mismatch with phenotype. J. Antimicrob. Chemother. 2021, 76, 2215–2217. [Google Scholar] [CrossRef] [PubMed]
  146. McMurtry, T.A.; Barekat, A.; Rodriguez, F.; Purewal, P.; Bulman, Z.P.; Lenhard, J.R. Capability of Enterococcus faecalis to shield Gram-negative pathogens from aminoglycoside exposure. J. Antimicrob. Chemother. 2021, 76, 2610–2614. [Google Scholar] [CrossRef] [PubMed]
  147. Galimand, M.; Schmitt, E.; Panvert, M.; Desmolaize, B.; Douthwaite, S.; Mechulam, Y.; Courvalin, P. Intrinsic resistance to aminoglycosides in Enterococcus faecium is conferred by the 16S rRNA m5C1404-specific methyltransferase EfmM. RNA 2011, 17, 251–262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Eliopoulos, G.M.; Farber, B.F.; Murray, B.E.; Wennersten, C.; Moellering, R.C. Ribosomal resistance of clinical enterococcal to streptomycin isolates. Antimicrob. Agents Chemother. 1984, 25, 398–399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Lerma, L.L.; Benomar, N.; Valenzuela, A.S.; del Muñoz, M.C.C.; Gálvez, A.; Abriouel, H. Role of EfrAB efflux pump in biocide tolerance and antibiotic resistance of Enterococcus faecalis and Enterococcus faecium isolated from traditional fermented foods and the effect of EDTA as EfrAB inhibitor. Food Microbiol. 2014, 44, 249–257. [Google Scholar] [CrossRef]
  150. Bender, J.K.; Cattoir, V.; Hegstad, K.; Sadowy, E.; Coque, T.M.; Westh, H.; Hammerum, A.M.; Schaffer, K.; Burns, K.; Murchan, S.; et al. Update on prevalence and mechanisms of resistance to linezolid, tigecycline and daptomycin in enterococci in Europe: Towards a common nomenclature. Drug Resist. Updates 2018, 40, 25–39. [Google Scholar] [CrossRef]
  151. Courvalin, P. Vancomycin resistance in Gram-positive cocci. Clin. Infect. Dis. 2006, 42, 25–34. [Google Scholar] [CrossRef]
  152. Xu, X.; Lin, D.; Yan, G.; Ye, X.; Wu, S.; Guo, Y.; Zhu, D.; Hu, F.; Zhang, Y.; Wang, F.; et al. vanM, a new glycopeptide resistance gene cluster found in Enterococcus faecium. Antimicrob. Agents Chemother. 2010, 54, 4643–4647. [Google Scholar] [CrossRef] [Green Version]
  153. Ahmed, M.O.; Baptiste, K.E. Vancomycin-resistant enterococci: A review of antimicrobial resistance mechanisms and perspectives of human and animal health. Microb. Drug Resist. 2018, 24, 590–606. [Google Scholar] [CrossRef] [Green Version]
  154. Mühlberg, E.; Umstätter, F.; Kleist, C.; Domhan, C.; Mier, W.; Uhl, P. Renaissance of vancomycin: Approaches for breaking antibiotic resistance in multidrug-resistant bacteria. Can. J. Microbiol. 2020, 66, 11–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Boyd, D.A.; Willey, B.M.; Fawcett, D.; Gillani, N.; Mulvey, M.R. Molecular characterization of Enterococcus faecalis N06-0364 with low-level vancomycin resistance harboring a novel D-Ala-D-Ser gene cluster, vanL. Antimicrob. Agents Chemother. 2008, 52, 2667–2672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Lebreton, F.; Depardieu, F.; Bourdon, N.; Fines-Guyon, M.; Berger, P.; Camiade, S.; Leclercq, R.; Courvalin, P.; Cattoir, V. D-Ala-D-Ser VanN-type transferable vancomycin resistance in Enterococcus faecium. Antimicrob. Agents Chemother. 2011, 55, 4606–4612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Sassi, M.; Guérin, F.; Lesec, L.; Isnard, C.; Fines-Guyon, M.; Cattoir, V.; Giard, J.C. Genetic characterization of a VanG-type vancomycin-resistant Enterococcus faecium clinical isolate. J. Antimicrob. Chemother. 2018, 73, 852–855. [Google Scholar] [CrossRef]
  158. Arthur, M.; Depardieu, F.; Gerbaud, G.; Galimand, M.; Leclercq, R.; Courvalin, P. The VanS sensor negatively controls VanR-mediated transcriptional activation of glycopeptide resistance genes of Tn1546 and related elements in the absence of induction. J. Bacteriol. 1997, 179, 97–106. [Google Scholar] [CrossRef] [Green Version]
  159. Arthur, M.; Molinas, C.; Courvalin, P. The VanS-VanR two-component regulatory system controls synthesis of depsipeptide peptidoglycan precursors in Enterococcus faecium BM4147. J. Bacteriol. 1992, 174, 2582–2591. [Google Scholar] [CrossRef] [Green Version]
  160. Arthur, M.; Molinas, C.; Courvalin, P. Sequence of the vanY gene required for production of a vancomycin-inducible D,D-carboxypeptidase in Enterococcus faecium BM4147. Gene 1992, 120, 111–114. [Google Scholar] [CrossRef]
  161. Reynolds, P.E.; Depardieu, F.; Dutka-Malen, S.; Arthur, M.; Courvalin, P. Glycopeptide resistance mediated by enterococcal transposon Tn1546 requires production of VanX for hydrolysis of D-alanyl-D-alanine. Mol. Microbiol. 1994, 13, 1065–1070. [Google Scholar] [CrossRef]
  162. Arthur, M.; Depardieu, F.; Molinas, C.; Reynolds, P.; Courvalin, P. The vanZ gene of Tn1546 from Enterococcus faecium BM4147 confers resistance to teicoplanin. Gene 1995, 154, 87–92. [Google Scholar] [CrossRef]
  163. Leclercq, R.; Dutka-Malen, S.; Duval, J.; Courvalin, P. Vancomycin resistance gene vanC is specific to Enterococcus gallinarum. Antimicrob. Agents Chemother. 1992, 36, 2005–2008. [Google Scholar] [CrossRef] [Green Version]
  164. Navarro, F.; Courvalin, P. Analysis of genes encoding D-alanine-D-alanine ligase-related enzymes in Enterococcus casseliflavus and Enterococcus flavescens. Antimicrob. Agents Chemother. 1994, 38, 1788–1793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Stogios, P.J.; Savchenko, A. Molecular mechanisms of vancomycin resistance. Protein Sci. 2020, 29, 654–669. [Google Scholar] [CrossRef] [PubMed]
  166. Hawkey, P.M. Mechanisms of quinolone action and microbial response. J. Antimicrob. Chemother. 2003, 51, 29–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Arsène, S.; Leclercq, R. Role of a qnr-like, gene in the intrinsic resistance of Enterococcus faecalis to fluoroquinolones. Antimicrob. Agents Chemother. 2007, 51, 3254–3258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Nakanishi, N.; Yoshida, S.; Wakebe, H.; Inoue, M.; Mitsuhashi, S. Mechanisms of clinical resistance to fluoroquinolones in Enterococcus faecalis. Antimicrob. Agents Chemother. 1991, 35, 1053–1059. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Korten, V.; Huang, M.U.N.; Murray, B.E. Analysis by PCR and direct DNA sequencing of gyrA mutations associated with fluoroquinolone resistance in Enterococcus faecalis. Antimicrob. Agents Chemother. 1994, 38, 2091–2094. [Google Scholar] [CrossRef] [Green Version]
  170. Tankovic, J.; Mahjoubi, F.; Courvalin, P.; Duval, J.; Leclercq, R. Development of fluoroquinolone resistance in Enterococcus faecalis and role of mutations in the DNA Gyrase gyrA Gene. Antimicrob. Agents Chemother. 1996, 40, 2558–2561. [Google Scholar] [CrossRef] [Green Version]
  171. Kanematsu, E.; Deguchi, T.; Yasuda, M.; Kawamura, T.; Nishino, Y.; Kawada, Y. Alterations in the GyrA subunit of DNA gyrase and the ParC subunit of DNA topoisomerase IV associated with quinolone resistance in Enterococcus faecalis. Antimicrob. Agents Chemother. 1998, 42, 433–435. [Google Scholar] [CrossRef] [Green Version]
  172. el Amin, N.; Jalal, S.; Wretlind, B. Alterations in gyrA and parC associated with fluoroquinolone resistance in Enterococcus faecium. Antimicrob. Agents Chemother. 1999, 43, 947–949. [Google Scholar] [CrossRef] [Green Version]
  173. López, M.; Tenorio, C.; del Campo, R.; Zarazaga, M.; Torres, C. Characterization of the mechanisms of fluoroquinolone resistance in vancomycin-resistant enterococci of different origins. J. Chemother. 2011, 23, 87–91. [Google Scholar] [CrossRef]
  174. Yasufuku, T.; Shigemura, K.; Shirakawa, T.; Matsumoto, M.; Nakano, Y.; Tanaka, K.; Arakawa, S.; Kawabata, M.; Fujisawa, M. Mechanisms of and risk factors for fluoroquinolone resistance in clinical Enterococcus faecalis isolates from patients with urinary tract infections. J. Clin. Microbiol. 2011, 49, 3912–3916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. bin Kim, Y.; Seo, H.J.; Seo, K.W.; Jeon, H.Y.; Kim, D.K.; Kim, S.W.; Lim, S.K.; Lee, Y.J. Characteristics of high-level ciprofloxacin-resistant Enterococcus faecalis and Enterococcus faecium from retail chicken meat in Korea. J. Food Prot. 2018, 81, 1357–1363. [Google Scholar]
  176. Yamanaka, H.; Kadomatsu, R.; Takagi, T.; Ohsawa, M.; Yamamoto, N.; Kubo, N.; Takemoto, T.; Ohsawa, K. Antimicrobial resistance profiles of vancomycin-resistant Enterococcus species isolated from laboratory mice. J. Vet. Sci. 2019, 20, e13. [Google Scholar] [CrossRef]
  177. Oyamada, Y.; Ito, H.; Fujimoto, K.; Asada, R.; Niga, T.; Okamoto, R.; Inoue, M.; Yamagishi, J.I. Combination of known and unknown mechanisms confers high-level resistance to fluoroquinolones in Enterococcus faecium. J. Med. Microbiol. 2006, 55, 729–736. [Google Scholar] [CrossRef]
  178. Hooper, D.C. Fluoroquinolone resistance among Gram-positive cocci. Lancet Infect. Dis. 2002, 2, 530–538. [Google Scholar] [CrossRef]
  179. Jonas, B.M.; Murray, B.E.; Weinstock, G.M. Characterization of emeA, a norA homolog and multidrug resistance efflux pump, in Enterococcus faecalis. Antimicrob. Agents Chemother. 2001, 45, 3574–3579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Tran, J.H.; Jacoby, G.A. Mechanism of plasmid-mediated quinolone resistance. Proc. Natl. Acad. Sci. USA 2002, 99, 5638–5642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Tran, J.H.; Jacoby, G.A.; Hooper, D.C. Interaction of the plasmid-encoded quinolone resistance protein Qnr with Escherichia coli DNA gyrase. Antimicrob. Agents Chemother. 2005, 49, 118–125. [Google Scholar] [CrossRef] [Green Version]
  182. Tran, J.H.; Jacoby, G.A.; Hooper, D.C. Interaction of the plasmid-encoded quinolone resistance protein QnrA with Escherichia coli Topoisomerase IV. Antimicrob. Agents Chemother. 2005, 49, 3050–3052. [Google Scholar] [CrossRef] [Green Version]
  183. Agersø, Y.; Pedersen, A.G.; Aarestrup, F.M. Identification of Tn5397-like and Tn916-like transposons and diversity of the tetracycline resistance gene tet(M) in enterococci from humans, pigs and poultry. J. Antimicrob. Chemother. 2006, 57, 832–839. [Google Scholar] [CrossRef] [Green Version]
  184. Cauwerts, K.; Decostere, A.; de Graef, E.M.; Haesebrouck, F.; Pasmans, F. High prevalence of tetracycline resistance in Enterococcus isolates from broilers carrying the erm(B) gene. Avian Pathol. 2007, 36, 395–399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Chotinantakul, K.; Chansiw, N.; Okada, S. Antimicrobial resistance of Enterococcus spp. isolated from Thai fermented pork in Chiang Rai Province, Thailand. J. Glob. Antimicrob. Resist. 2018, 12, 143–148. [Google Scholar] [CrossRef] [PubMed]
  186. Woźniak-Biel, A.; Bugla-Płoskońska, G.; Burdzy, J.; Korzekwa, K.; Ploch, S.; Wieliczko, A. Antimicrobial resistance and biofilm formation in Enterococcus spp. isolated from humans and turkeys in Poland. Microb. Drug Resist. 2019, 25, 277–286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Rice, L.B. Tn916 family conjugative transposons and dissemination of antimicrobial resistance determinants. Antimicrob. Agents Chemother. 1998, 42, 1871–1877. [Google Scholar] [CrossRef] [Green Version]
  188. Sharkey, L.K.R.; Edwards, T.A.; O’Neill, A.J. ABC-F proteins mediate antibiotic resistance through ribosomal protection. MBio 2016, 7, e01975-15. [Google Scholar] [CrossRef] [Green Version]
  189. Antonelli, A.; D’Andrea, M.M.; Brenciani, A.; Galeotti, C.L.; Morroni, G.; Pollini, S.; Varaldo, P.E.; Rossolini, G.M. Characterization of poxtA, a novel phenicol-oxazolidinone-tetracycline resistance gene from an MRSA of clinical origin. J. Antimicrob. Chemother. 2018, 73, 1763–1769. [Google Scholar] [CrossRef] [Green Version]
  190. Brenciani, A.; Fioriti, S.; Morroni, G.; Cucco, L.; Morelli, A.; Pezzotti, G.; Paniccià, M.; Antonelli, A.; Magistrali, C.F.; Rossolini, G.M.; et al. Detection in Italy of a porcine Enterococcus faecium isolate carrying the novel phenicol-oxazolidinone-tetracycline resistance gene poxtA. J. Antimicrob. Chemother. 2019, 74, 817–818. [Google Scholar] [CrossRef]
  191. Lei, C.W.; Kang, Z.Z.; Wu, S.K.; Chen, Y.P.; Kong, L.H.; Wang, H.N. Detection of the phenicol–oxazolidinone–tetracycline resistance gene poxtA in Enterococcus faecium and Enterococcus faecalis of food-producing animal origin in China. J. Antimicrob. Chemother. 2019, 74, 2459–2461. [Google Scholar] [CrossRef]
  192. Fioriti, S.; Coccitto, S.N.; Cedraro, N.; Simoni, S.; Morroni, G.; Brenciani, A.; Mangiaterra, G.; Vignaroli, C.; Vezzulli, L.; Biavasco, F.; et al. Linezolid resistance genes in enterococci isolated from sediment and zooplankton in two Italian coastal areas. Appl. Environ. Microbiol. 2021, 87, e02958-20. [Google Scholar] [CrossRef]
  193. Wang, J.L.; Hsueh, P.R. Therapeutic options for infections due to vancomycin-resistant enterococci. Expert Opin. Pharmacother. 2009, 10, 785–796. [Google Scholar] [CrossRef]
  194. Lee, B.J.; Vu, B.N.; Seddon, A.N.; Hodgson, H.A.; Wang, S.K. Treatment considerations for CNS infections caused by vancomycin-resistant Enterococcus faecium: A focused review of linezolid and daptomycin. Ann. Pharmacother. 2020, 54, 1243–1251. [Google Scholar] [CrossRef] [PubMed]
  195. Riccardi, N.; Monticelli, J.; Antonello, R.M.; di Lallo, G.; Frezza, D.; Luzzati, R.; di Bella, S. Therapeutic options for infections due to vanB genotype vancomycin-resistant enterococci. Microb. Drug Resist. 2021, 27, 536–545. [Google Scholar] [CrossRef]
  196. Wingler, M.J.; Patel, N.R.; King, S.T.; Wagner, J.L.; Barber, K.E.; Stover, K.R. Linezolid for the treatment of urinary tract infections caused by vancomycin-resistant enterococci. Pharmacy 2021, 9, 175. [Google Scholar] [CrossRef] [PubMed]
  197. Wang, Y.; Lv, Y.; Cai, J.; Schwarz, S.; Cui, L.; Hu, Z.; Zhang, R.; Li, J.; Zhao, Q.; He, T.; et al. A novel gene, optrA, that confers transferable resistance to oxazolidinones and phenicols and its presence in Enterococcus faecalis and Enterococcus faecium of human and animal origin. J. Antimicrob. Chemother. 2015, 70, 2182–2190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Bi, R.; Qin, T.; Fan, W.; Ma, P.; Gu, B. The emerging problem of linezolid-resistant enterococci. J. Glob. Antimicrob. Resist. 2018, 13, 11–19. [Google Scholar] [CrossRef]
  199. Càmara, J.; Camoez, M.; Tubau, F.; Pujol, M.; Ayats, J.; Ardanuy, C.; Domínguez, M.Á. Detection of the novel optrA gene among linezolid-resistant enterococci in Barcelona, Spain. Microb. Drug Resist. 2019, 25, 87–93. [Google Scholar] [CrossRef] [Green Version]
  200. Elghaieb, H.; Freitas, A.R.; Abbassi, M.S.; Novais, C.; Zouari, M.; Hassen, A.; Peixe, L. Dispersal of linezolid-resistant enterococci carrying poxtA or optrA in retail meat and food-producing animals from Tunisia. J. Antimicrob. Chemother. 2019, 74, 2865–2869. [Google Scholar] [CrossRef] [PubMed]
  201. Ruiz-Ripa, L.; Feßler, A.T.; Hanke, D.; Eichhorn, I.; Azcona-Gutiérrez, J.M.; Pérez-Moreno, M.O.; Seral, C.; Aspiroz, C.; Alonso, C.A.; Torres, L.; et al. Mechanisms of linezolid resistance among enterococci of clinical origin in Spain—Detection of optrA- and cfr(D)-carrying E. faecalis. Microorganisms 2020, 8, 1155. [Google Scholar] [CrossRef]
  202. Freitas, A.R.; Finisterra, L.; Tedim, A.P.; Duarte, B.; Novais, C.; Peixe, L. Linezolid- and multidrug-resistant enterococci in raw commercial dog food, Europe, 2019–2020. Emerg. Infect. Dis. 2021, 27, 2221–2224. [Google Scholar] [CrossRef] [PubMed]
  203. Kloss, P.; Xiong, L.; Shinabarger, D.L.; Mankin, A.S. Resistance mutations in 23S rRNA identify the site of action of the protein synthesis inhibitor linezolid in the ribosomal peptidyl transferase center. J. Mol. Biol. 1999, 294, 93–101. [Google Scholar] [CrossRef] [Green Version]
  204. Mališová, L.; Jakubů, V.; Pomorská, K.; Musílek, M.; Žemličková, H. Spread of linezolid-resistant Enterococcus spp. in human clinical isolates in the Czech Republic. Antibiotics 2021, 10, 219. [Google Scholar] [CrossRef] [PubMed]
  205. Marshall, S.H.; Donskey, C.J.; Hutton-Thomas, R.; Salata, R.A.; Rice, L.B. Gene dosage and linezolid resistance in Enterococcus faecium and Enterococcus faecalis. Antimicrob. Agents Chemother. 2002, 46, 3334–3336. [Google Scholar] [CrossRef] [Green Version]
  206. Kehrenberg, C.; Schwarz, S.; Jacobsen, L.; Hansen, L.H.; Vester, B. A new mechanism for chloramphenicol, florfenicol and clindamycin resistance: Methylation of 23S ribosomal RNA at A2503. Mol. Microbiol. 2005, 57, 1064–1073. [Google Scholar] [CrossRef]
  207. Long, K.S.; Poehlsgaard, J.; Kehrenberg, C.; Schwarz, S.; Vester, B. The Cfr rRNA methyltransferase confers resistance to phenicols, lincosamides, oxazolidinones, pleuromutilins, and streptogramin A antibiotics. Antimicrob. Agents Chemother. 2006, 50, 2500–2505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Deshpande, L.M.; Ashcraft, D.S.; Kahn, H.P.; Pankey, G.; Jones, R.N.; Farrell, D.J.; Mendes, R.E. Detection of a new cfr-like gene, cfr(B), in Enterococcus faecium isolates recovered from human specimens in the United States as part of the SENTRY Antimicrobial Surveillance Program. Antimicrob. Agents Chemother. 2015, 59, 6256–6261. [Google Scholar] [CrossRef] [Green Version]
  209. Bender, J.K.; Fleige, C.; Klare, I.; Fiedler, S.; Mischnik, A.; Mutters, N.T.; Dingle, K.E.; Werner, G. Detection of a cfr(B) variant in German Enterococcus faecium clinical isolates and the impact on linezolid resistance in Enterococcus spp. PLoS ONE 2016, 11, e0167042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Pang, S.; Boan, P.; Lee, T.; Gangatharan, S.; Tan, S.J.; Daley, D.; Lee, Y.T.; Coombs, G.W. Linezolid-resistant ST872 Enteroccocus faecium harbouring optrA and cfr(D) oxazolidinone resistance genes. Int. J. Antimicrob. Agents 2019, 55, 105831. [Google Scholar] [CrossRef] [PubMed]
  211. Guerin, F.; Sassi, M.; Dejoies, L.; Zouari, A.; Schutz, S.; Potrel, S.; Auzou, M.; Collet, A.; Lecointe, D.; Auger, G.; et al. Molecular and functional analysis of the novel cfr(D) linezolid resistance gene identified in Enterococcus faecium. J. Antimicrob. Chemother. 2020, 75, 1699–1703. [Google Scholar] [CrossRef] [PubMed]
  212. Cui, L.; Wang, Y.; Lv, Y.; Wang, S.; Song, Y.; Li, Y.; Liu, J.; Xue, F.; Yang, W.; Zhang, J. Nationwide surveillance of novel oxazolidinone resistance gene optrA in Enterococcus isolates in China from 2004 to 2014. Antimicrob. Agents Chemother. 2016, 60, 7490–7493. [Google Scholar] [CrossRef] [Green Version]
  213. Li, P.; Yang, Y.; Ding, L.; Xu, X.; Lin, D. Molecular investigations of linezolid resistance in enterococci optrA variants from a hospital in Shanghai. Infect. Drug Resist. 2020, 13, 2711–2716. [Google Scholar] [CrossRef]
  214. Regulation (UE), No. 528/2012 of 22 of May 2012; Official Journal of the European Union—L167/1; European Parliament and Council: Brussels, Belgium, 2012.
  215. Maillard, J.-Y. Resistance of bacteria to biocides. Microbiol. Spectr. 2018, 6, 109–126. [Google Scholar] [CrossRef] [PubMed]
  216. Roedel, A.; Dieckmann, R.; Brendebach, H.; Hammerl, J.A.; Kleta, S.; Noll, M.; al Dahouk, S.; Vinczea, S. Biocide-tolerant Listeria monocytogenes isolates from German food production plants do not show cross-resistance to clinically relevant antibiotics. Appl. Environ. Microbiol. 2019, 85, e01253-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Maillard, J.-Y. Factors Affecting the Activities of Microbicides. In Russell, Hugo & Ayliffe’s Principles and Practice of Disinfection, Preservation & Sterilization, 5th ed.; Fraise, A.P., Lambert, P.A., Maillard, J.-Y., Eds.; Blackwell Publishing: London, UK, 2013; pp. 71–86. [Google Scholar]
  218. Sinel, C.; Augagneur, Y.; Sassi, M.; Bronsard, J.; Cacaci, M.; Guérin, F.; Sanguinetti, M.; Meignen, P.; Cattoir, V.; Felden, B. Small RNAs in vancomycin-resistant Enterococcus faecium involved in daptomycin response and resistance. Sci. Rep. 2017, 7, 11067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Michaux, C.; Hartke, A.; Martini, C.; Reiss, S.; Albrecht, D.; Budin-Verneuil, A.; Sanguinetti, M.; Engelmann, S.; Hain, T.; Verneuil, N.; et al. Involvement of Enterococcus faecalis small RNAs in stress response. Infect. Immun. 2014, 82, 3599–3611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Shioya, K.; Michaux, C.; Kuenne, C.; Hain, T.; Verneuil, N.; Hartsch, T.; Hartke, A.; Giard, J. Genome-wide identification of small RNAs in the opportunistic pathogen Enterococcus faecalis V583. PLoS ONE 2017, 6, e23948. [Google Scholar] [CrossRef] [PubMed]
  221. Dejoies, L.; le Neindre, K.; Reissier, S.; Felden, B.; Cattoir, V. Distinct expression profiles of regulatory RNAs in the response to biocides in Staphylococcus aureus and Enterococcus faecium. Sci. Rep. 2021, 11, 6892. [Google Scholar] [CrossRef] [PubMed]
  222. Peng, Z.; Ehrmann, M.A.; Waldhuber, A.; Niemeyer, C.; Miethke, T.; Frick, J.; Xiong, T.; Vogel, R.F. Phosphotransferase systems in Enterococcus faecalis OG1RF enhance anti- stress capacity in vitro and in vivo. Res. Microbiol. 2017, 168, 558–566. [Google Scholar] [CrossRef]
  223. Rizzotti, L.; Rossi, F.; Torriani, S. Biocide and antibiotic resistance of Enterococcus faecalis and Enterococcus faecium isolated from the swine meat chain. Food Microbiol. 2016, 60, 160–164. [Google Scholar] [CrossRef]
  224. Suchomel, M.; Lenhardt, A.; Kampf, G.; Grisold, A. Enterococcus hirae, Enterococcus faecium and Enterococcus faecalis show different sensitivities to typical biocidal agents used for disinfection. J. Hosp. Infect. 2019, 103, 435–440. [Google Scholar] [CrossRef]
  225. Schwaiger, K.; Harms, K.S.; Bischoff, M.; Preikschat, P.; Mölle, G.; Bauer-Unkauf, I.; Lindorfer, S.; Thalhammer, S.; Bauer, J.; Hölzel, C.S. Insusceptibility to disinfectants in bacteria from animals, food and humans—Is there a link to antimicrobial resistance? Front. Microbiol. 2014, 5, 88. [Google Scholar] [CrossRef] [Green Version]
  226. Wieland, N.; Boss, J.; Lettmann, S.; Fritz, B.; Schwaiger, K.; Bauer, J.; Hölzel, C.S. Susceptibility to disinfectants in antimicrobial-resistant and -susceptible isolates of Escherichia coli, Enterococcus faecalis and Enterococcus faecium from poultry–ESBL/AmpC-phenotype of E. coli is not associated with resistance to a quaternary ammonium. J. Appl. Microbiol. 2017, 122, 1508–1517. [Google Scholar] [CrossRef] [PubMed]
  227. Esmaeilishirazifard, E.; Usher, L.; Trim, C.; Denise, H.; Sangal, V.; Tyson, G.H.; Barlow, A.; Redway, K.F.; Taylor, J.D.; Kremyda-Vlachou, M.; et al. Bacterial adaptation to venom in snakes and arachnida. Microbiol. Spectr. 2022, e02408-21. [Google Scholar] [CrossRef] [PubMed]
  228. Bengoechea, J.A.; Bamford, C.G.G. SARS-CoV-2, bacterial co-infections, and AMR: The deadly trio in COVID-19? EMBO Mol. Med. 2020, 12, e12560. [Google Scholar] [CrossRef]
  229. DeVoe, C.; Segal, M.R.; Wang, L.; Stanley, K.; Madera, S.; Fan, J.; Schouest, J.; Graham-Ojo, R.; Nichols, A.; Prasad, P.A.; et al. Increased rates of secondary bacterial infections, including Enterococcus bacteremia, in patients hospitalized with coronavirus disease 2019 (COVID-19). Infect. Control Epidemiol. 2021, 1–8. [Google Scholar] [CrossRef] [PubMed]
  230. Toc, D.A.; Mihaila, R.M.; Botan, A.; Bobohalma, C.N.; Risteiu, G.A.; Simut-Cacuci, B.N.; Steorobelea, B.; Troanca, S.; Junie, L.M. Enterococcus and COVID-19: The Emergence of a Perfect Storm? Int. J. Transl. Med. 2022, 2, 220–229. [Google Scholar] [CrossRef]
  231. Kampmeier, S.; Tönnies, H.; Correa-Martinez, C.L.; Mellmann, A.; Schwierzeck, V. A nosocomial cluster of vancomycin resistant enterococci among COVID-19 patients in an intensive care unit. Antimicrob. Resist. Infect. Control 2020, 9, 154. [Google Scholar] [CrossRef]
  232. Gaibani, P.; D’Amico, F.; Bartoletti, M.; Lombardo, D.; Rampelli, S.; Fornaro, G.; Coladonato, S.; Siniscalchi, A.; Re, M.C.; Viale, P.; et al. The Gut Microbiota of Critically Ill Patients With COVID-19. Front. Cell. Infect. Microbiol. 2021, 11, 670424. [Google Scholar] [CrossRef]
  233. Toc, D.A.; Butiuc-Keul, A.L.; Iordache, D.; Botan, A.; Mihaila, R.M.; Costache, C.A.; Colosi, I.A.; Chiorean, C.; Neagoe, D.S.; Gheorghiu, L.; et al. Descriptive Analysis of Circulating Antimicrobial Resistance Genes in Vancomycin-Resistant Enterococcus (VRE) during the COVID-19 Pandemic. Biomedicines 2022, 10, 1122. [Google Scholar] [CrossRef]
  234. Jeon, K.; Jeong, S.; Lee, N.; Park, M.-J.; Song, W.; Kim, H.-S.; Kim, H.S.; Kim, J.-S. Impact of COVID-19 on Antimicrobial Consumption and Spread of Multidrug-Resistance in Bacterial Infections. Antibiotics 2022, 11, 535. [Google Scholar] [CrossRef]
  235. Bienzle, D.; Rousseau, J.; Marom, D.; MacNicol, J.; Jacobson, L.; Sparling, S.; Prystajecky, N.; Fraser, E.; Weese, J.S. Risk Factors for SARS-CoV-2 Infection and Illness in Cats and Dogs. Emerg. Infect. Dis. 2022, 28, 1154–1162. [Google Scholar] [CrossRef]
Figure 1. Summary diagram of the most relevant virulence factors present in enterococci.
Figure 1. Summary diagram of the most relevant virulence factors present in enterococci.
Antibiotics 11 00857 g001
Figure 2. Representation of the organization of the vanA and vanB operons. PR, PH, PY: promoting regions. Adapted from Refs. [8,151,153].
Figure 2. Representation of the organization of the vanA and vanB operons. PR, PH, PY: promoting regions. Adapted from Refs. [8,151,153].
Antibiotics 11 00857 g002
Table 1. Representation of all of the aminoglycoside-modifying enzymes that have been described in the Enterococcus genus, along with the type of resistance presented and the respective reference.
Table 1. Representation of all of the aminoglycoside-modifying enzymes that have been described in the Enterococcus genus, along with the type of resistance presented and the respective reference.
EnzymeType of Resistance ConferredReference
AACsAAC(6′)-IiIntrinsicLow- to moderate-level resistance to tobramycin and kanamycin[125,126]
AAC(6′)-Id[127]
AAC(6′)-Ih[127]
APHsAPH(3′)-IIIaExtrinsicLow-level resistance to kanamycin and amikacin[128]
APH(2″)-IbExtrinsicHigh-level resistance to gentamicin[129]
APH(2″)-Ic[130]
APH(2″)-Id[131]
APH(2″)-Ie[132]
ANTsANT(6′)-IaExtrinsicHigh-level resistance to streptomycin[3,133]
ANT(3″)-Ia or ANT(3″)(9)[3]
ANT(4″)-Ia[134]
Bifunctional Enzyme (AAC + APH)AAC-6′-Ie-APH-2ExtrinsicHigh-level resistance to gentamicin[135,136,137]
Table 2. Summary table of the main mechanisms of resistance and decreased susceptibility to antibiotics and biocides, presented by the Enterococcus genus.
Table 2. Summary table of the main mechanisms of resistance and decreased susceptibility to antibiotics and biocides, presented by the Enterococcus genus.
Antibiotics
Group of AntibioticsResistance TypeMechanism of ResistanceAssociated Genes
β-LactamIntrinsicLow affinity PBPs that do not allow for easy antibiotic bindingpbp5/pbp4
AcquiredMutations that lead to alteration in PBPs’ molecular structure and cause an even lower affinity-
AminoglycosidesIntrinsicPoor antibiotic uptake trough the cell wall-
IntrinsicModification of the antibiotic moleculeaac
AcquiredModification of the antibiotic moleculeaph, ant, aac-aph
IntrinsicTarget-site modification through rRNA methyltransferaseefmM
AcquiredTarget-site modification through point mutations-
-Efflux of the antibioticefrAB
GlycopeptidesIntrinsic/AcquiredTarget-site modificationvan operons
FluoroquinolonesAcquiredTarget-site modification through gene mutationgyrA, pacC
-Efflux of the antibioticemeA
-Target-site protectionqnrE.faecalis
Tetracyclines-Efflux of the antibiotictet(K), tet(L)
-Target-site protectiontet(M), tet(O), tet(S)
Intrinsic/AcquiredTarget-site protectionpoxtA
OxazolidinonesAcquiredTarget-site modification through point mutations-
Intrinsic/AcquiredTarget-site protectionpoxtA
AcquiredTarget-site protectionoptrA
Biocides
sRNAsBacterial survival in stressful environmental conditions such as in the presence of biocides
PTS Systems
ChtRS
Efflux PumpsQacA/B and EfrAB—efflux of different biocides
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Geraldes, C.; Tavares, L.; Gil, S.; Oliveira, M. Enterococcus Virulence and Resistant Traits Associated with Its Permanence in the Hospital Environment. Antibiotics 2022, 11, 857. https://doi.org/10.3390/antibiotics11070857

AMA Style

Geraldes C, Tavares L, Gil S, Oliveira M. Enterococcus Virulence and Resistant Traits Associated with Its Permanence in the Hospital Environment. Antibiotics. 2022; 11(7):857. https://doi.org/10.3390/antibiotics11070857

Chicago/Turabian Style

Geraldes, Catarina, Luís Tavares, Solange Gil, and Manuela Oliveira. 2022. "Enterococcus Virulence and Resistant Traits Associated with Its Permanence in the Hospital Environment" Antibiotics 11, no. 7: 857. https://doi.org/10.3390/antibiotics11070857

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

Article Metrics

Back to TopTop