Next Article in Journal
Factors Associated with Progression of Atrial Fibrillation and Impact on All-Cause Mortality in a Cohort of European Patients
Next Article in Special Issue
Effect of Different Sealers on the Cytocompatibility and Osteogenic Potential of Human Periodontal Ligament Stem Cells: An In Vitro Study
Previous Article in Journal
Mitochondria: Emerging Consequential in Sickle Cell Disease
Previous Article in Special Issue
Benefits and Challenges of the Use of Two Novel vB_Efa29212_2e and vB_Efa29212_3e Bacteriophages in Biocontrol of the Root Canal Enterococcus faecalis Infections
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Regulations of Essential WalRK Two-Component System on Enterococcus faecalis

1
Orthopedic Research Institute, Department of Orthopedics, West China Hospital, Sichuan University, Chengdu 610017, China
2
West China School of Public Health and West China Fourth Hospital, Sichuan University, Chengdu 610017, China
3
West China Hospital of Stomatology, Sichuan University, Chengdu 610017, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
J. Clin. Med. 2023, 12(3), 767; https://doi.org/10.3390/jcm12030767
Submission received: 9 November 2022 / Revised: 4 January 2023 / Accepted: 10 January 2023 / Published: 18 January 2023
(This article belongs to the Special Issue Clinical Updates in Endodontics)

Abstract

:
Enterococcus faecalis (E. faecalis) is a Gram-positive, facultative anaerobic bacterium that is highly adaptable to its environment. In humans, it can cause serious infections with biofilm formation. With increasing attention on its health threat, prevention and control of biofilm formation in E. faecalis have been observed. Many factors including polysaccharides as well as autolysis, proteases, and eDNA regulate biofilm formation. Those contributors are regulated by several important regulatory systems involving the two-component signal transduction system (TCS) for its adaptation to the environment. Highly conserved WalRK as one of 17 TCSs is the only essential TCS in E. faecalis. In addition to biofilm formation, various metabolisms, including cell wall construction, drug resistance, as well as interactions among regulatory systems and resistance to the host immune system, can be modulated by the WalRK system. Therefore, WalRK has been identified as a key target for E. faecalis infection control. In the present review, the regulation of WalRK on E. faecalis pathogenesis and associated therapeutic strategies are demonstrated.

Graphical Abstract

1. Introduction

Enterococcus faecalis (E. faecalis) is a Gram-positive, facultative anaerobic bacterium, which almost serves as normal colonizing bacteria in the gastrointestinal tracts of humans and is generally related to commensal life in the gastrointestinal tract [1]. E. faecalis has strong environmental adaptability. The abilities to form robust and mature biofilms [2,3,4] makes E. faecalis able to survive in scarce nutrition and resist extreme alkaline pH values [5,6,7,8]. With biofilm formation, E. faecalis can also cause serious infection. As one of the most common pathogens in hospital infections, E. faecalis leads to various opportunistic infections [9,10,11,12,13] such as urinary tract infections, bacteremia, prosthetic joint infections, abdominal-pelvic infections, endocarditis [14,15,16], periodontal disease, and endodontic infections [17,18]. Despite its pathogenicity, E. faecalis is also a probiotic and is often found as part of the autochthonous flora in a variety of fermented foods, namely cheeses and sausages [19,20,21,22,23].
Among the regulatory systems in E. faecalis [24], the two-component signal transduction system (TCS) is one of the signal pathways necessary for bacteria to adapt to the environmental changes, which helps bacteria to perceive and respond to external signals rapidly. A TCS consists of two basic components: histidine kinase (HK), which is anchored to the cell membrane, and intracellular response regulators (RRs). The HK responds to the environmental stimulus by auto-phosphorylating on a conserved histidine residue in the cytoplasmic domain. Then, the phosphate group is transferred to an aspartate residue on the partner RR. Through affinity for target promoter DNA, RR effects a corresponding modulation of transcription and elicits a cellular response to the environmental stimuli [25,26].
E. faecalis includes 17 members of two-component systems, in which the CroRS system is required for intrinsic β-lactam resistance [27,28]. The β-lactams are structural analogs of the D-Ala4-D-Ala5, an extremity of peptidoglycan precursors, and act as suicide substrates of the D, D-transpeptidases that catalyze the last cross-linking step of peptidoglycan synthesis [29]. The CroS HK activates in response to the cell wall damage induced by antimicrobial effects, and CroR upregulates an alternative penicillin-binding protein (PBP), carrying out transpeptidase activity with a low-binding affinity for β-lactams [30]. By enhancing the release of nucleic acids into the biofilm matrix, the two-component Fsr-QS (quorum sensing) system controls the production of extracellular gelatinase and contributes to the biofilm formation [31]. The fsr regulatory locus is comprised of three genes, designated as FsrA, FsrB, and FsrC, which are necessary for positive regulation of the virulence-associated proteases Gelatinase E (GelE) and serine protease (SprE) [32,33]. FsrC is a histidine kinase that senses extracellular accumulation of a peptide lactone encoded at the C-terminus of the FsrB protein. The sensing of FsrC leads to activation of the response regulator and transcription factor FsrA [34]. In addition, the Fsr system positively regulates degradation of important immunopeptides, exerts an inhibitory effect against the complement system, and performs a critical role in overcoming the immune systems that are inherent to human serum [34,35,36]. Glycopeptide antibiotics vancomycin (VM) and teicoplanin (TE) inhibit the extracellular steps of bacterial peptidoglycan synthesis by binding to the C-terminal D-alanyl-D-alanine (D-Ala-D-Ala) residues of cell wall precursors. Synthesis of modified peptidoglycan precursors terminating in D-lactate, glycopeptide such as peptidoglycan precursor UDP-N-acetylmuramyl-L-Ala-γ-D-Glu-L-Lys-D-Ala-D-lactate (UDP-MurNAc-pentadepsipeptide) makes a more than 1000-fold decreased affinity for glycopeptides resistance. In E. faecalis, the TCS VanSRB regulates an immediately downstream promoter and contributes to VanB-type vancomycin resistance. Vancomycin is identified as an inducer for the VanSBRB activation [27,37]. Among all 17 TCSs, walRK is the only essential TCS for E. faecalis, which determines the viability of stains, indicating that the survival of bacteria can be significantly affected by the knockout of walRK [26].
The WalRK (also named YycFG or VicRK) TCS signaling pathway is one of the most widely distributed TCS systems, highly conserved in most low-G+C Gram-positive bacteria [38]. The system has been reported as necessary for several closely related pathogens including Staphylococcus aureus [39], Streptococcus pneumoniae [40], and Bacillus subtilis [41]. The core of the WalRK TCS consists of HK (WalK) on the cell surface and an intracellular RR (WalR) [25]. The WalK histidine kinases contain HAMP (a domain present in histidine kinases, adenylyl cyclases, methyl-accepting proteins, and phosphatases) and PAS (Per-Arnt-Sim) sensing domains, which promote two-way conformational communication between the input and output domains signaling proteins, along with histidine phosphotransfer (HisKA) and kinase catalytic (HATPase) domains [42,43,44,45]. The amino acid sequences of the receiver and effector domains of WalR are highly conserved and belong to the OmpR family of response regulators [44,45,46]. In addition to walR and walK, yycH, yycI, and yycJ are also included in the E. faecalis WalRK TCS [47,48]. The auxiliary proteins YycHI play roles in signaling through the WalK [45]. Although the cytoplasmic, conserved YycJ auxiliary protein contains a putative metal binding site in a β-lactamase fold [49], its functions are currently unknown [45]. In this review, we investigate the role of the WalRK system in the metabolism of E. faecalis, including biofilm formation, cell wall construction, drug resistance, and resistance to the host immune system. In addition, potentially therapeutic strategies associated with the WalRK TCS are noted for clinical infections administration.

2. Material and Methods

Based on searching the PubMed database with terms “WalR, WalK, two-component system, Enterococcus faecalis, regulation, and treatment”, we conduct a literature search. The inclusion criteria included: (1) original studies or reviews on related topics; (2) published in English; (3) full-text. The exclusion criteria included: (1) abstracts from conferences and (2) document cannot be downloaded. The search in PubMed yielded 426 articles and, after exclusion, 122 were included in our literature manager. To supply our possibly ignored information, associated references in the already included literature were also detected.

3. WalKR Two-Component Signal Transduction System and Resistance of Enterococcus Faecalis

3.1. Regulation of E. faecalis on Biofilm Formation

E. faecalis is characterized by its ability to survive in a high-sodium-chloride concertation and tolerate both acid and alkaline conditions. With the widespread use of antibiotics, E. faecalis as a pathogenic bacterium has gradually emerged with antibiotic resistance. Those adaptive capabilities are closely related to the E. faecalis biofilm formation.
Biofilm is mainly developed by packed microcolonies and extracellular polymeric substances (EPSs). Multiple genes have been identified to play an important role in biofilm biogenesis. Genes encode proteinaceous adhesins such as LPxTG surface proteins, autolysins, and glycolipids, contributing to intercellular adhesion in biofilm formation. EF3314 is a putative surface-exposed antigenic protein in E. faecalis. It is considered as an adhesin for biofilm formation.
Polysaccharides, as well as autolysins and proteases, are thought to contribute to EPS production, and therefore affect biofilm maturation. Gene cluster epa (enterococcal polysaccharide antigen) is involved in the biosynthesis of cell-wall-associated polysaccharides. In E. faecalis, the epa gene cluster encodes enzymes and transporters involving synthesis of nucleotide sugar precursors in the cytoplasm, formation and polymerization of repeating units, and export to the cell surface, which contribute to biofilm formation, resistance to enterococcal polysaccharide antigen (PMN) killing, and virulence [50].
eDNA is an important component of the extracellular matrix of bacterial biofilms, providing structural stability to the biofilm and protection against antimicrobials. Murein hydrolases, also referred to as autolysins, have been implicated in biofilm production [51]. In E. faecalis, autolysins such as AtlA, AtlB, and AtlC are involved in eDNA release. AtlA is identified as an N-acetylglucosaminidase [52]. The synergistical action with two additional peptidoglycan hydrolases AtlB and AtlC DNA regulates the release of eDNA [53]. The deletion of atlA not only delays the biofilm formation but also inhibits cells division and reduces the rate of cellular lysis [53].
Via proteolytic processing of the accumulation-associated protein, bacterial proteases contribute to biofilm formation [54,55]. Two proteases produced by E. faecalis such as GelE and SprE have been shown to regulate bacterial autolysis as well as eDNA release and, thus, contribute to E. faecalis biofilm formation, playing important roles in the pathogenesis of various diseases, including endophthalmitis, peritonitis, endocarditis, and orthopedic implant infections [56]. Regarding orthopedic implant infections, E. faecalis with high gelatinase production is a prevalent isolate [57]. The quorum-sensing Fsr two-component system is comprised of three genes involving fsrA, fsrB, and fsrC. By sensing the extracellular accumulation of a peptide lactone encoded at the C terminus of the FsrB protein, the FsrC histidine kinase is activated, subsequently irrigating response regulator FsrA. Inactivation of the fsr-controlled gene gelE inhibits the biofilm formation [31].
In enterococcal surface proteins (Esps), LPxTG-type surface proteins can covalently immobilize the surface protein to the cell-wall peptidoglycan, involved in cell-to-cell adhesion and biofilm formation [58]. LPxTG-type surface proteins may be partially mediated by the N-terminal moiety of the protein and interacts with additional E. faecalis-specific factors to enhance the biofilm formation [58,59].
Lipoteichoic acid (LTA) and polysaccharide are also involved in biofilm formation via interactions with environmental molecules. In E. faecalis, the dltABCD operon is required to obtain D-alanylation of LTA, which is an essential constituent of the Gram-positive bacterial cell wall [60]. The deletion of the dltA gene induces the absence of D-alanine in the LTA and a stronger negative net charge on the bacterial cell surface. D-alanine esters as positively charged groups can modulate negatively charged teichoic acids. Subsequently, the biofilm formation was significantly decreased in E. faecalis with reduced adherence to epithelial cells, while the susceptibility to cationic antimicrobial peptides was increased [61,62].
Enterococcal binding substance (EBS), a kind of cognate receptor for the aggregation substance, such as cell wall LTA, is necessary for mating complexes in biofilm formation [59,63]. Aggregation substances (ASs), such as Asa1, Asp1, and Acs10, promote the conjugation by directing bacterial aggregation, enables cell–cell contact between donor and recipient strains, and increases the E. faecalis virulence [64,65]. In a dose-dependent manner, Asc10 with an LTA aggregation domain is located near the N-terminal of E. Faecalis LTA [66]. The AS protein is comprised of an N-terminal domain (output signal), a variable region, a central domain (responsible for aggregation), and two Arg-Gly-Asp (RGD) motifs (implicate in binding host cell integrins) [66,67,68]. The expression of the AS protein can be induced by the peptide pheromone, which is secreted from plasmid-free recipient cells, or can be induced by host factors that may act by shifting the effective ratios of endogenous cCF10/iCF10 [65,69,70,71,72,73,74].
The E. faecalis WalRK system regulates the various metabolisms including cell wall construction, osmotic protection, biofilm formation, and drug resistance, which are significantly related to bacterial pathogenicity [39]. Furthermore, the drug resistance in E. faecalis infection is worrisome, and even some Enterococcus spp. are completely resistant to all antimicrobial drug regimens [12,75]. The regulation of E. faecalis by the WalRK system includes the physiological metabolisms, biofilm and cell wall formation, drug resistance, and immunity reaction with the host. Considering the WalRK regulations above, it as a novel target can be developed and available for infection disease, especially for the multidrug-resistant vancomycin-resistant Enterococci (VRE).
Histidine kinase WalK is a transmembrane protein that senses external signals and possesses kinase, phosphotransferase, and phosphatase activities, while the response regulator WalR contains a functional domain (REC) that accepts a phosphate group for autophosphorylation. The kinase ATP-binding domain (HATPase_c) of WalK is stimulated and activated by extracellular signals and then causes phosphorylation at the histidine site of WalK (HisKA domain). The phosphoryl group is first transmitted from the phosphorylated WalK and then to the aspartic acid residue of WalR, leading to conformational change and exposure of the DNA-binding site. Phosphorylated WalR then binds the DNA promoter motif and is involved in transcriptional regulation of bacteria metabolism [48]. In our previous study [76], an antisense walR negatively regulates the walR expression, leading to the biofilm formation reduction and extracellular polysaccharide synthesis, significantly decreasing the pathogenicity of E. faecalis. The roles of WalRK in biofilm formation attracts increasing interest on a potential target of antibacterial and antibiofilm [77]. However, the mechanism of WalRK modulating biofilm formation remains largely unrevealed [78].

3.2. Antibiotic Resistance of Enterococcus faecalis

Antibiotic-resistant E. faecalis is of concern and contributes to difficult-to-treat nosocomial infections [79]. The main reasons for multidrug-resistant enterococci emergence include intrinsic resistance to antimicrobial agents such as beta lactams and aminoglycosides, and mobile elements acquired resistance against glycopeptides, quinolones, tetracyclines, and streptogramin [80]. The acquired vanA and vanB genotypes are the most common glycopeptide resistance in E. faecalis and are regulated by the VanRS and VanRBSB TCSs [81]. The vancomycin resistance operon modifies the vancomycin-binding target by synthesis of peptidoglycan precursors that terminate in D-lactate or D-serine. Instead of normal D-alanine sidechain terminus, the modified vancomycin resistance operon significantly decreases the affinity vancomycin on E. faecalis [82]. With the evolution of resistance, the last-resort antibiotics, such as daptomycin and linezolid, for multidrug-resistant Enterococcal infections are limited [79].
E. faecalis produces a specific penicillin-binding protein (PBP5) that mediates high-level intrinsic resistance to the cephalosporin class of β-lactam antibiotics. Comenge et al. first described that the regulatory CroRS TCS is essential for PBP5-mediated β-lactam resistance [17]. In further dissection of the CroRS TCS, a noncognate histidine kinase CisS is capable of influencing CroR to mediate resistance to cell wall stresses such as cephalosporins and glycopeptides. The cross-talk of CroRS and CisRS responds to the vancomycin [83]. Vancomycin is the only identified stimuli for CisRS thus far and CisRS can also interact with VanRS for resistance [84].
Daptomycin (DAP) is a lipopeptide antibiotic with potent in vitro bactericidal activity against VRE [85]. By binding with calcium, DAP can form a cationic moiety and disrupts membranes to play its potential antimicrobial role [86]. However, reports of DAP-insensitive enterococci strains have become increasingly common [86]. Tran et al. identified the resistance of Enterococcus faecium (E. Faecium) to daptomycin associated with YycFG system mutation. In addition, this resistance mechanism was also related with the LiaFSR system in E. faecium [49]. The LiaFSR three-component system controls the cell envelope stress response of E. Faecalis and its mutation results in the resistance of daptomycin [87]. E. Faecalis and E. Faecium both belong to Enterococcus, and they share a great similarity in gene composition. From this point, it may be inferred that the YycFG system of E. Faecalis also plays a role in DAP resistance.
The two-component systems in bacteria can cross-talk with each other to regulate a variety of cellular processes, including biofilm formation, virulence expression, and drug resistance [26]. For Gram-positive bacteria, the relationship between the WalRK TCS and other TCSs such as VanSB, CovRS, and SaeSR has been explored. In Streptococcus mutans (S. mutans), VicR and CovR can both regulate a group of genes of extracellular polysaccharides synthesis. However, VicR positively regulates these genes in most cases, while CovR modulates inversely [88]. For S. aureus, the VicRK TCS can activate the SaeSR TCS to regulate virulence factors [89]. In E. faecalis, VanSB can phosphorylate VicR from S. pneumoniae and raise a species cross-talking [6,26]. However, there is rare evidence suggesting interaction between WalRK and other TCSs such as VanSR in E. faecalis.

3.3. Regulatory Roles of the WalRK TCS in Response to Host Immunity

TCSs in pathogenic bacteria can also adjust responsible immunity activities and infect microenvironments in hosts. For S. aureus, WalRK positively regulates virulence genes through a virulence factor regulation system SaeSR TCS, promoting the polymorphonuclear leukocytes (PMNs) lysis and playing an essential role in innate immunity evasion [90]. Additionally, adaptive immunity can enhance the effect of innate immune cells and influence host susceptibility to microorganism invaders [91]. By modulating the virulence expression, the evasion mechanisms from the adaptive immune system are developed [89,92]. In S. aureus, a second immunoglobulin binding protein (Sbi) is secreted as a virulence factor, binding with the Fc portion of IgG, and interferes with host adaptive immune responses. In a previous study, WalRK positively regulated biofilms aggregation and virulence genes expression, controlling the progress of the infection condition in the host [93]. However, the specific mechanism of the WalRK TCS on modulating the interaction of the host and E. faecalis needs to be further explored.

4. Chemical Compounds Act on WalRK to Inhibit the Pathogenicity of Bacteria

Due to the important role of WalRK in Gram-positive bacteria such as E. faecalis, S. aureus, S. epidermidis, and S. mutans [94,95,96,97], many attempts were applied on WalRK to inhibit pathogenicity [94].
Clinically, iodoform is widely used as a disinfectant. It has a strong effect on aerobic Gram-positive bacteria and can enhance antibacterial activity against E. faecalis when combined with calcium hydroxide [98]. Therefore, it is used in dental treatment, and can be used for the treatment of periapical periodontitis when in combination with calcium hydroxide [99]. However, because of its tendency to cause allergic reactions and cytotoxicity to epithelial cells and macrophages, new drugs need to be developed [100].

4.1. Antibiotics-Based Agents on WalRK

Many studies are committed to looking for antibacterial substances from microorganism, such as actinomycin, a good natural antibiotic derived from Actinomycetes.
The autophosphorylation of WalK is an essential step for bacterial pathogenicity. An actinomycin metabolite, waldiomycin, belongs to the family of anguclin antibiotics, which is similar to dioxin in structure, and it can affect the expression of related genes on cell wall metabolism, cell proliferation, and aggregation. Specifically, this inhibitor was found to play a role in inhibiting the function of WalK phosphorylation, attributed to the antibacterial effects mentioned above [94,96]. Subsequently, this waldiomycin has an antibacterial activity against Gram-positive bacteria, including methicillin-resistant S. aureus (MRSA) and B. subtilis. Okada et al. pointed out that by binding to the WalK cytoplasmic domain, walkmycin B, which was the main product of Streptomyces sp. (MK632-100F11), inhibited WalK phosphorylation [101]. The H-box region is an essential motif of WalK, and Kato et al. found that waldiomycin interfered with various cellular processes of B. subtilis by targeting the H-box region of WalK and other HK [102].
Zerumbone1 is a kind of monocyclic sesquiterpene whose derivatives have high biological activities, including its ring opening derivative 2 (Figure 1). Kitayama et al. pointed out that Zerumbone ring opening derivative 2 can inhibit the autophosphorylation of WalK in B. subtilis, while it cannot inhibit drug-resistant bacteria, such as MRSA and VRE [103].
At the same time, signermycin B was found in a temperature-sensitive B. subtilis walR mutant (CNM2000) [104]. The signermycin B can inhibit self-phosphorylation by targeting the conserved dimer domain of WalK and inhibit the activity of WalR to hinder the function of the walKR system, achieving an antibacterial effect.
As the existence of E. faecalis biofilm has greatly improved the resistance of E. faecalis to existing antibiotic therapy, Lin et al. investigated the effect of derivatives from a kind of probiotic named B. subtilis natto on the E. faecalis biofilm formation [105]. Natto treatment showed downregulation of the E. faecalis walRK system transcription, which resulted in peptidoglycan and glycolipid biosynthesis inhibition and a reduction in extracellular polysaccharide synthesis. These studies showed that B. subtilis natto may help to control the biofilm of E. faecalis, and then affect the drug resistance of E. faecalis.

4.2. Novel Screened Antibacterial Substances on WalRK

Recently, many novel antibacterial substances have been developed via high-throughput screening systems with a high efficiency.
Qin et al. verified a series of novel inhibitors on the S. epidermidis WalK protein by using structure-based virtual screening (SBVs) in a small-molecule lead compound library [106]. There, seven inhibitors belonging to four different chemical structures were identified: three thiazolidinone analogs (compounds 2, 5, and 7), two benzamide analogs (compounds 1 and 3), a furan derivative (compound 4), and a pyrimidine ketone derivative (compound 6). By replacing different functional groups in Compound 2 (C27H28O5N2S) (Figure 2), a series of derivatives were designed, such as H2-38, H2-39, H2-74, and H2-81 (Table 1) [107]. Those derivatives as new antibacterial agents targeting WalK can effectively eradicate against clinical Staphylococcus isolates such as multidrug-resistant Staphylococcus spp. and S. aureus. H2-60 (C27H16F2O2N3S2) and H2-81(C27H17FO2N3S2) are two thiazolidinone derivatives. Both derivatives are proved to have an effect on WalK. Biofilm has greatly improved the resistance of E. faecalis to existing antibiotic therapy. In E. faecalis, Chen et al. identified two derivatives that play a role against biofilm production by repression of WalK kinase phosphorylation activity [108]. When combined with traditional antibiotics such as daptomycin, H2-60 can even synergistically enhance the sensitivity of E. faecalis than single application.
In other species, such as Streptococcus agalactiae, the verified H2-60 and H2-81 can significantly inhibit biofilm formation and bacterial growth [109]. Furuta et al. screened and identified a compound I-8-15 (1-dodecyl-2-isopropyl imidazole), which can specifically inhibit the WalK of S. aureus, resulting in conformational changes and inhibiting dimerization of WalK [110]. By this mechanism, I-8-15 showed a significant antibacterial activity against MRSA and VRE.
In addition, Gotoh et al. also identified wallycin A (4-methoxy-1-naphthol) and wallycin B (1,6-dimethyl-3-[4-(trifluoromethyl) phenyl] pyrimido [5,4-e][1,2,4]triazine-5, 7-dione) compounds through the high-throughput screening system [95]. IclR is a regulatory protein containing the N-terminal DNA binding domain of the wing helix trans helix motif and the C-terminal regulatory domain of the protein, and the binding of signal molecules to the C-terminal domain of IclR is considered to inhibit the transcription of target genes by regulating DNA binding activity or the polymerization of IclR. By analysis, it was proved that these two compounds can inhibit the phosphate transfer from P-WalKtru to WalR. As a result, they can inactivate the IclR-WalR chimeric repressor, as walR inhibitors.

4.3. Nucleic-Acid-Based Interfering walRK for Treatment

Antisense RNA (asRNA) is a single-stranded RNA reverse complement to the targeted mRNA. It can inhibit the translation or transcription of target mRNA by recruiting the RNase to degrade RNA duplex structures. Using this property, RNA interference technology is developed to inhibit the transcription of target genes to affect physiological functions [111].
E. faecalis is the main pathogen of persistent periapical periodontitis. The E. faecalis biofilm formation is related to the WalRK modulation. How to inhibit WalRK expression and reduce biofilm formation attracts a prospective research topic. Our previous study identified an endogenous antisense walR (ASwalR) inhibiting walR expression [112]. With ASwalR interference, the biofilm formation of E. faecalis is significantly reduced and sensitivity to traditional chlorhexidine is increased. To further improve the efficiency of ASwalR, a graphene-oxide (GO)–polyethyleneimine (PEI)-based transformation strategy is applied, indicating that it can be used as a potential supplementary therapy for infection treatment [113].

4.4. Bacteriophage Therapy

Phage is the general name of viruses that infect microorganisms such as bacteria, fungi, algae, actinomycetes, and spirochetes. The size of phages is generally small, with an incomplete cell structure, and only contains a single nucleic acid. In addition, some of the phages can cause the lysis of host bacteria. By nucleic acid invasion and replication, the phages can kill bacteria as a novel treatment mode. Unlike traditional antibiotics, phages replicate themselves in the bacteria, and this repeating process effectively amplifies its antibacterial efficiencies. Because of antibiotic resistance and the fact that antibiotics may destroy the microbial balance in the body, phage therapy is less likely to develop resistance and is also highly microbial-specific [114].
However, phage therapy has some corresponding disadvantages including the specificity to target bacteria, easily producing corresponding tolerance under selection, being contaminated by bacteria used for culture, and rapidly being removed from the body and other unclear efficacies [115]. Therefore, many studies have been performed to improve the above shortcomings.
To overcome short duration, Merril et al. screened a mutation of the main phage head protein E in capsid proteins of phage K and phage P22, in which a long-half-life phage was acquired [116]. Meanwhile, chemical modification with polyethylene glycol (PEG) can significantly improve the cycle half-life of phages by reducing the release level of helper-T-cell-type-Ⅰ-related cytokines (IFN-γ and IL-6) [117].
The biofilm formation significantly decreases the sensitivity of bacteria on phages. To strengthen the effect of phages, Lu modified a phage to encode enzymes [118]. Enzymes can hydrolyze extracellular polymeric substances and help the phage to bind to the receptor. Antibiotic efflux pumps are the essential resistance mechanisms. The combination of antibiotic treatment and bacteriophages acting on the expression of antibiotic efflux pumps can make bacteria more sensitive to antibiotics [119,120].
In E. faecalis, phages can potentially be a genetic scheme for treatment. Duerkop et al. found sewage phages in urban sewage [121]. Through phage infection proteins for Enterococcus faecalis (PIPEF), sewage phages inhibit the colonization of E. faecalis in the intestine, playing a therapeutic role.
Therefore, the phage therapy is also expected as a good scheme for the infection treatment. We proposed a hypothetical treatment, which integrates the antisense fragment of the walRK gene or inhibitor encoding the WalRK protein on the phage genome specific to E. faecalis and modified by PEG, and then injects it into the human body to treat E. faecalis infection to inhibit the biofilm formation, cooperating with corresponding antibiotic treatment.

5. Conclusions

Biofilm formation in E. faecalis induces a high adaption to the environment and causes severe infections. Many factors including polysaccharides as well as autolysis, proteases, and eDNA regulate biofilm formation. These factors are regulated by several important regulatory systems involving the TCSs. The highly conserved WalRK, one of 17 TCSs, is the only essential TCS in E. faecalis. In addition, various metabolisms, including cell wall construction, drug resistance, and interactions between regulatory systems and resistance to the host immune system, can be regulated by the WalRK system. Therefore, WalRK has been identified as a key target for the infection control. At present, novel antibiotics, screened antibacterial substances, and nucleic-acid-based agents interfering on WalRK phosphorylation or expression have been investigated for infection treatment. Meanwhile, promising phage therapy, a potentially genetic scheme, shows a good inhibition on bacterial growth. Therefore, suppressing the action of WalRK on bacterial infection with programed phages such as antisense fragment integration will be a topic of great interest.

Author Contributions

Conceptualization, Formal Analysis, S.W. and Q.P.; Data Collection, J.Z. and R.F.; Writing—Original Draft Preparation, J.Z. and R.F.; Writing—Review and Editing, S.W., Q.P. and L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

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. Facklam, R. What happened to the streptococci: Overview of taxonomic and nomenclature changes. Clin. Microbiol. Rev. 2002, 15, 613–630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Rams, T.E.; Feik, D.; Young, V.; Hammond, B.F.; Slots, J. Enterococci in human periodontitis. Oral Microbiol. Immunol. 1992, 7, 249–252. [Google Scholar] [CrossRef] [PubMed]
  3. Distel, J.W.; Hatton, J.F.; Gillespie, M.J. Biofilm formation in medicated root canals. J. Endod. 2002, 28, 689–693. [Google Scholar] [CrossRef] [PubMed]
  4. Mohamed, J.A.; Huang, D.B. Biofilm formation by enterococci. J. Med. Microbiol. 2007, 56 Pt 2, 1581–1588. [Google Scholar] [CrossRef] [Green Version]
  5. Figdor, D.; Davies, J.K.; Sundqvist, G. Starvation survival, growth and recovery of Enterococcus faecalis in human serum. Oral Microbiol. Immunol. 2003, 18, 234–239. [Google Scholar] [CrossRef]
  6. McHugh, C.P.; Zhang, P.; Michalek, S.; Eleazer, P.D. pH required to kill Enterococcus faecalis in vitro. J. Endod. 2004, 30, 218–219. [Google Scholar] [CrossRef] [PubMed]
  7. Sedgley, C.M.; Lennan, S.L.; Appelbe, O.K. Survival of Enterococcus faecalis in root canals ex vivo. Int. Endod. J. 2005, 38, 735–742. [Google Scholar] [CrossRef] [Green Version]
  8. Stevens, R.H.; Grossman, L.I. Evaluation of the antimicrobial potential of calcium hydroxide as an intracanal medicament. J. Endod. 1983, 9, 372–374. [Google Scholar] [CrossRef]
  9. Gilmore, M.S.; Ferretti, J.J. Microbiology. The thin line between gut commensal and pathogen. Science 2003, 299, 1999–2002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Hunt, C.P. The emergence of enterococci as a cause of nosocomial infection. Br. J. Biomed. Sci. 1998, 55, 149–156. [Google Scholar]
  11. Jett, B.D.; Huycke, M.M.; Gilmore, M.S. Virulence of enterococci. Clin. Microbiol. Rev. 1994, 7, 462–478. [Google Scholar] [CrossRef] [PubMed]
  12. Moellering, R.C., Jr. Emergence of Enterococcus as a significant pathogen. Clin. infect. Dis. 1992, 14, 1173–1176. [Google Scholar] [CrossRef] [PubMed]
  13. Woodford, N. Glycopeptide-resistant enterococci: A decade of experience. J. Med. Microbiol. 1998, 47, 849–862. [Google Scholar] [CrossRef]
  14. Ali, L.; Goraya, M.U.; Arafat, Y.; Ajmal, M.; Chen, J.L.; Yu, D. Molecular Mechanism of Quorum-Sensing in Enterococcus faecalis: Its Role in Virulence and Therapeutic Approaches. Int. J. Mol. Sci. 2017, 18, 960. [Google Scholar] [CrossRef] [Green Version]
  15. Marino, A.; Munafò, A.; Zagami, A.; Ceccarelli, M.; Di Mauro, R.; Cantarella, G.; Bernardini, R.; Nunnari, G.; Cacopardo, B. Ampicillin Plus Ceftriaxone Regimen against Enterococcus faecalis Endocarditis: A Literature Review. J. Clin. Med. 2021, 10, 4594. [Google Scholar] [CrossRef] [PubMed]
  16. Romay, E.; Pericàs, J.M.; García-País, M.J.; Hernández-Meneses, M.; Ayuso, B.; García-González, J.; Garcés-Durán, R.V.; Rabuñal, R.; Alonso-García, P.; García-Garrote, F.; et al. On Behalf Of Lucus Augusti And Hospital Clinic Endocarditis Teams. Relationship among Streptococcus gallolyticus Subsp. gallolyticus, Enterococcus faecalis and Colorectal Neoplasms in Recurrent Endocarditis: A Historical Case Series. J. Clin. Med. 2022, 11, 2181. [Google Scholar] [CrossRef] [PubMed]
  17. Comenge, Y.; Quintiliani, R., Jr.; Li, L.; Dubost, L.; Brouard, J.P.; Hugonnet, J.E.; Arthur, M. The CroRS two-component regulatory system is required for intrinsic beta-lactam resistance in Enterococcus faecalis. J. Bacteriol. 2003, 185, 7184–7192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Frère, J.M.; Joris, B. Penicillin-sensitive enzymes in peptidoglycan biosynthesis. Crit. Rev. Microbiol. 1985, 11, 299–396. [Google Scholar] [CrossRef]
  19. Tinoco, J.M.; Buttaro, B.; Zhang, H.; Liss, N.; Sassone, L.; Stevens, R. Effect of a genetically engineered bacteriophage on Enterococcus faecalis biofilms. Arch. Oral Biol. 2016, 71, 80–86. [Google Scholar] [CrossRef] [Green Version]
  20. Lindenstrauss, A.G.; Behr, J.; Ehrmann, M.A.; Haller, D.; Vogel, R.F. Identification of fitness determinants in Enterococcus faecalis by differential proteomics. Arch. Microbiol. 2013, 195, 121–130. [Google Scholar] [CrossRef]
  21. Centeno, J.A.; Menendez, S.; Hermida, M.; Rodríguez-Otero, J.L. Effects of the addition of Enterococcus faecalis in Cebreiro cheese manufacture. Int. J. Food Microbiol. 1999, 48, 97–111. [Google Scholar] [CrossRef] [PubMed]
  22. Nieto-Arribas, P.; Seseña, S.; Poveda, J.M.; Chicón, R.; Cabezas, L.; Palop, L. Enterococcus populations in artisanal Manchego cheese: Biodiversity, technological and safety aspects. Food Microbiol. 2011, 28, 891–899. [Google Scholar] [CrossRef] [PubMed]
  23. Metaxopoulos, J.; Samelis, J.; Papadelli, M. Technological and Microbiological Evaluation of Traditional Processes as Modified for the Industrial Manufacturing of Dry Fermented Sausage in Greece. Ital. J. Food Sci. 2001, 13, 3–18. [Google Scholar]
  24. Marchesini, B.; Bruttin, A.; Romailler, N.; Moreton, R.S.; Stucchi, C.; Sozzi, T. Microbiological events during commercial meat fermentations. J. Appl. Bacteriol. 1992, 73, 203–209. [Google Scholar] [CrossRef] [PubMed]
  25. Wunderlich, P.F.; Braun, L.; Fumagalli, I.; D’Apuzzo, V.; Heim, F.; Karly, M.; Lodi, R.; Politta, G.; Vonbank, F.; Zeltner, L. Double-blind report on the efficacy of lactic acid-producing Enterococcus SF68 in the prevention of antibiotic-associated diarrhoea and in the treatment of acute diarrhoea. J. Int. Med. Res. 1989, 17, 333–338. [Google Scholar] [CrossRef]
  26. Hancock, L.E.; Perego, M. Systematic inactivation and phenotypic characterization of two-component signal transduction systems of Enterococcus faecalis V583. J. Bacteriol. 2004, 186, 7951–7958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Ma, P.; Yuille, H.M.; Blessie, V.; Göhring, N.; Iglói, Z.; Nishiguchi, K.; Nakayama, J.; Henderson, P.J.; Phillips-Jones, M.K. Expression, purification and activities of the entire family of intact membrane sensor kinases from Enterococcus faecalis. Mol. Membr. Biol. 2008, 25, 449–473. [Google Scholar] [CrossRef] [PubMed]
  28. Laguri, C.; Phillips-Jones, M.K.; Williamson, M.P. Solution structure and DNA binding of the effector domain from the global regulator PrrA (RegA) from Rhodobacter sphaeroides: Insights into DNA binding specificity. Nucleic Acids Res. 2003, 31, 6778–6787. [Google Scholar] [CrossRef] [Green Version]
  29. Hoch, J.A.; Silhavy, T.J. Two-Component Signal Transduction; ASM Press: Washington, DC, USA, 1995; Volume xvi, p. 488. [Google Scholar]
  30. Tierney, A.R.; Rather, P.N. Roles of two-component regulatory systems in antibiotic resistance. Future Microbiol. 2019, 14, 533–552. [Google Scholar] [CrossRef]
  31. Hancock, L.E.; Perego, M. The Enterococcus faecalis fsr two-component system controls biofilm development through production of gelatinase. J. Bacteriol. 2004, 186, 5629–5639. [Google Scholar] [CrossRef] [Green Version]
  32. Qin, X.; Singh, K.V.; 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] [PubMed] [Green Version]
  33. Sifri, C.D.; Mylonakis, E.; Singh, K.V.; Qin, X.; Garsin, D.A.; Murray, B.E.; Ausubel, F.M.; Calderwood, S.B. Virulence effect of Enterococcus faecalis protease genes and the quorum-sensing locus fsr in Caenorhabditis elegans and mice. infect. Immun. 2002, 70, 5647–5650. [Google Scholar] [CrossRef] [PubMed]
  34. Goh, H.M.S.; Yong, M.H.A.; Chong, K.K.L.; Kline, K.A. Model systems for the study of Enterococcal colonization and infection. Virulence 2017, 8, 1525–1562. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. 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] [PubMed] [Green Version]
  36. 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] [PubMed] [Green Version]
  37. Evers, S.; Courvalin, P. Regulation of VanB-type vancomycin resistance gene expression by the VanS(B)-VanR (B) two-component regulatory system in Enterococcus faecalis V583. J. Bacteriol. 1996, 178, 1302–1309. [Google Scholar] [CrossRef] [Green Version]
  38. Hancock, L.; Perego, M. Two-component signal transduction in Enterococcus faecalis. J. Bacteriol. 2002, 184, 5819–5825. [Google Scholar] [CrossRef] [Green Version]
  39. Kuroda, M.; Kuwahara-Arai, K.; Hiramatsu, K. Identification of the up- and down-regulated genes in vancomycin-resistant Staphylococcus aureus strains Mu3 and Mu50 by cDNA differential hybridization method. Biochem. Biophys. Res. Commun. 2000, 269, 485–490. [Google Scholar] [CrossRef]
  40. West, A.H.; Stock, A.M. Histidine kinases and response regulator proteins in two-component signaling systems. Trends Biochem. Sci. 2001, 26, 369–376. [Google Scholar] [CrossRef]
  41. Fabret, C.; Hoch, J.A. A two-component signal transduction system essential for growth of Bacillus subtilis: Implications for anti-infective therapy. J. Bacteriol. 1998, 180, 6375–6383. [Google Scholar] [CrossRef]
  42. Inouye, M.; Dutta, R. Histidine Kinases in Signal Transduction; Academic Press: Amsterdam, The Netherlands; Boston, MA, USA, 2003; Volume xviii, 520p. [Google Scholar] [CrossRef]
  43. Mascher, T.; Helmann, J.D.; Unden, G. Stimulus perception in bacterial signal-transducing histidine kinases. Microbiol. Mol. Biol. Rev. 2006, 70, 910–938. [Google Scholar] [CrossRef] [PubMed]
  44. Winkler, M.E.; Hoch, J.A. Essentiality, bypass, and targeting of the YycFG (VicRK) two-component regulatory system in gram-positive bacteria. J. Bacteriol. 2008, 190, 2645–2648. [Google Scholar] [CrossRef] [Green Version]
  45. Gao, R.; Mack, T.R.; Stock, A.M. Bacterial response regulators: Versatile regulatory strategies from common domains. Trends Biochem. Sci. 2007, 32, 225–234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Stock, A.M.; Robinson, V.L.; Goudreau, P.N. Two-component signal transduction. Annu. Rev Biochem. 2000, 69, 183–215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Dubrac, S.; Bisicchia, P.; Devine, K.M.; Msadek, T. A matter of life and death: Cell wall homeostasis and the WalKR (YycGF) essential signal transduction pathway. Mol. Microbiol. 2008, 70, 1307–1322. [Google Scholar] [CrossRef]
  48. Tran, T.T.; Panesso, D.; Gao, H.; Roh, J.H.; Munita, J.M.; Reyes, J.; Diaz, L.; Lobos, E.A.; Shamoo, Y.; Mishra, N.N.; et al. Whole-genome analysis of a daptomycin-susceptible enterococcus faecium strain and its daptomycin-resistant variant arising during therapy. Antimicrob. Agents Chemother. 2013, 57, 261–268. [Google Scholar] [CrossRef] [Green Version]
  49. Wagner, C.; Saizieu, A.d.; Schönfeld, H.J.; Kamber, M.; Lange, R.; Thompson, C.J.; Page, M.G. Genetic analysis and functional characterization of the Streptococcus pneumoniae vic operon. infect. Immun. 2002, 70, 6121–6128. [Google Scholar] [CrossRef] [Green Version]
  50. Teng, F.; Singh, K.V.; Bourgogne, A.; Zeng, J.; Murray, B.E. Further characterization of the epa gene cluster and Epa polysaccharides of Enterococcus faecalis. infect. Immun. 2009, 77, 3759–3767. [Google Scholar] [CrossRef] [Green Version]
  51. Rice, K.C.; Mann, E.E.; Endres, J.L.; Weiss, E.C.; Cassat, J.E.; Smeltzer, M.S.; Bayles, K.W. The cidA murein hydrolase regulator contributes to DNA release and biofilm development in Staphylococcus aureus. Proc. Natl. Acad. Sci. USA 2007, 104, 8113–8118. [Google Scholar] [CrossRef] [Green Version]
  52. Mesnage, S.; Chau, F.; Dubost, L.; Arthur, M. Role of N-acetylglucosaminidase and N-acetylmuramidase activities in Enterococcus faecalis peptidoglycan metabolism. J. Biol. Chem. 2008, 283, 19845–19853. [Google Scholar] [CrossRef] [Green Version]
  53. Guiton, P.S.; Hung, C.S.; Kline, K.A.; Roth, R.; Kau, A.L.; Hayes, E.; Heuser, J.; Dodson, K.W.; Caparon, M.G.; Hultgren, S.J. Contribution of autolysin and Sortase a during Enterococcus faecalis DNA-dependent biofilm development. infect. Immun. 2009, 77, 3626–3638. [Google Scholar] [CrossRef] [PubMed]
  54. Rohde, H.; Burdelski, C.; Bartscht, K.; Hussain, M.; Buck, F.; Horstkotte, M.A.; Knobloch, J.K.; Heilmann, C.; Herrmann, M.; Mack, D. Induction of Staphylococcus epidermidis biofilm formation via proteolytic processing of the accumulation-associated protein by staphylococcal and host proteases. Mol. Microbiol. 2005, 55, 1883–1895. [Google Scholar] [CrossRef] [PubMed]
  55. Martínez-García, S.; Rodríguez-Martínez, S.; Cancino-Diaz, M.E.; Cancino-Diaz, J.C. Extracellular proteases of Staphylococcus epidermidis: Roles as virulence factors and their participation in biofilm. APMIS 2018, 126, 177–185. [Google Scholar] [CrossRef]
  56. Nešuta, O.; Budešínský, M.; Hadravová, R.; Monincová, L.; Humpolicková, J.; Cerovský, V. How proteases from Enterococcus faecalis contribute to its resistance to short α-helical antimicrobial peptides. Pathog. Dis. 2017, 29, 75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Arciola, C.R.; Baldassarri, L.; Campoccia, D.; Creti, R.; Pirini, V.; Huebner, J.; Montanaro, L. Strong biofilm production, antibiotic multi-resistance and high gelE expression in epidemic clones of Enterococcus faecalis from orthopaedic implant infections. Biomaterials 2008, 29, 580–586. [Google Scholar] [CrossRef] [PubMed]
  58. Paganelli, F.L.; Willems, R.J.; Leavis, H.L. Optimizing future treatment of enterococcal infections: Attacking the biofilm? Trends Microbiol. 2012, 20, 40–49. [Google Scholar] [CrossRef]
  59. 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]
  60. Coupri, D.; Budin-Verneuil, A.; Hartke, A.; Benachour, A.; Léger, L.; Lequeux, T.; Pfund, E.; Verneuil, N. Genetic and pharmacological inactivation of d-alanylation of teichoic acids sensitizes pathogenic enterococci to β-lactams. J. Antimicrob. Chemother. 2019, 74, 3162–3169. [Google Scholar] [CrossRef]
  61. Chuang-Smith, O.N.; Wells, C.L.; Henry-Stanley, M.J.; Dunny, G.M. Acceleration of Enterococcus faecalis biofilm formation by aggregation substance expression in an ex vivo model of cardiac valve colonization. PLoS ONE 2010, 5, e15798. [Google Scholar] [CrossRef] [Green Version]
  62. Johnson, J.R.; Clabots, C.; Hirt, H.; Waters, C.; Dunny, G. Enterococcal aggregation substance and binding substance are not major contributors to urinary tract colonization by Enterococcus faecalis in a mouse model of ascending unobstructed urinary tract infection. infect. Immun. 2004, 72, 2445–2448. [Google Scholar] [CrossRef] [Green Version]
  63. Dunny, G.M. Enterococcal sex pheromones: Signaling, social behavior, and evolution. Annu. Rev. Genet. 2013, 47, 457–482. [Google Scholar] [CrossRef]
  64. Telford, J.L.; Barocchi, M.A.; Margarit, I.; Rappuoli, R.; Grandi, G. Pili in gram-positive pathogens. Nat. Rev. Microbiol. 2006, 4, 509–519. [Google Scholar] [CrossRef] [PubMed]
  65. Olmsted, S.B.; Kao, S.M.; van Putte, L.J.; Gallo, J.C.; Dunny, G.M. Role of the pheromone-inducible surface protein Asc10 in mating aggregate formation and conjugal transfer of the Enterococcus faecalis plasmid pCF10. J. Bacteriol. 1991, 173, 7665–7672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Waters, C.M.; Hirt, H.; McCormick, J.K.; Schlievert, P.M.; Wells, C.L.; Dunny, G.M. An amino-terminal domain of Enterococcus faecalis aggregation substance is required for aggregation, bacterial internalization by epithelial cells and binding to lipoteichoic acid. Mol. Microbiol. 2004, 52, 1159–1171. [Google Scholar] [CrossRef] [PubMed]
  67. Sava, I.G.; Heikens, E.; Huebner, J. Pathogenesis and immunity in enterococcal infections. Clin. Microbiol. Infect. 2010, 16, 533–540. [Google Scholar] [CrossRef] [PubMed]
  68. Schmitt, A.; Jiang, K.; Camacho, M.I.; Jonna, V.R.; Hofer, A.; Westerlund, F.; Christie, P.J.; Berntsson, R.P. PrgB promotes aggregation, biofilm formation, and conjugation through DNA binding and compaction. Mol. Microbiol. 2018, 109, 291–305. [Google Scholar] [CrossRef] [Green Version]
  69. Dunny, G.M.; Leonard, B.A.; Hedberg, P.J. Pheromone-inducible conjugation in Enterococcus faecalis: Interbacterial and host-parasite chemical communication. J. Bacteriol. 1995, 177, 871–876. [Google Scholar] [CrossRef] [Green Version]
  70. Leonard, B.A.; Podbielski, A.; Hedberg, P.J.; Dunny, G.M. Enterococcus faecalis pheromone binding protein, PrgZ, recruits a chromosomal oligopeptide permease system to import sex pheromone cCF10 for induction of conjugation. Proc. Natl. Acad. Sci. USA 1996, 93, 260–264. [Google Scholar] [CrossRef] [Green Version]
  71. Chandler, J.R.; Hirt, H.; Dunny, G.M. A paracrine peptide sex pheromone also acts as an autocrine signal to induce plasmid transfer and virulence factor expression in vivo. Proc. Natl. Acad. Sci. USA 2005, 102, 15617–15622. [Google Scholar] [CrossRef] [Green Version]
  72. Hirt, H.; Schlievert, P.M.; Dunny, G.M. In vivo induction of virulence and antibiotic resistance transfer in Enterococcus faecalis mediated by the sex pheromone-sensing system of pCF10. infect. Immun. 2002, 70, 716–723. [Google Scholar] [CrossRef] [Green Version]
  73. Galli, D.; Lottspeich, F.; Wirth, R. Sequence analysis of Enterococcus faecalis aggregation substance encoded by the sex pheromone plasmid pAD1. Mol. Microbiol. 1990, 4, 895–904. [Google Scholar] [CrossRef] [PubMed]
  74. Galli, D.; Friesenegger, A.; Wirth, R. Transcriptional control of sex-pheromone-inducible genes on plasmid pAD1 of Enterococcus faecalis and sequence analysis of a third structural gene for (pPD1-encoded) aggregation substance. Mol. Microbiol. 1992, 6, 1297–1308. [Google Scholar] [CrossRef] [PubMed]
  75. French, G.L. Enterococci and vancomycin resistance. Clin. Infect. Dis. 1998, 27 (Suppl. 1), S75–S83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Wu, S.; Liu, Y.; Zhang, H.; Lei, L. The Susceptibility to Calcium Hydroxide Modulated by the Essential walR Gene Reveals the Role for Enterococcus faecalis Biofilm Aggregation. J. Endod. 2019, 45, 295–301.e2. [Google Scholar] [CrossRef]
  77. Chen, Z.; Song, K.; Shang, Y.; Xiong, Y.; Lyu, Z.; Chen, J.; Zheng, J.; Li, P.; Wu, Y.; Gu, C.; et al. Selection and Identification of Novel Antibacterial Agents against Planktonic Growth and Biofilm Formation of Enterococcus faecalis. J. Med. Chem. 2021, 64, 15037–15052. [Google Scholar] [CrossRef]
  78. Rice, L.B. Enterococcal Physiology and Antimicrobial Resistance: The Streetlight Just Got a Little Brighter. mBio 2021, 12, e03511–e03520. [Google Scholar] [CrossRef]
  79. Van Harten, R.M.; Willems, R.J.L.; Martin, N.I.; Hendrickx, A.P.A. Multidrug-Resistant Enterococcal Infections: New Compounds, Novel Antimicrobial Therapies? Trends Microbiol. 2017, 25, 467–479. [Google Scholar] [CrossRef]
  80. Jahansepas, A.; Aghazadeh, M.; Rezaee, M.A.; Hasani, A.; Sharifi, Y.; Aghazadeh, T.; Mardaneh, J. Occurrence of Enterococcus faecalis and Enterococcus faecium in Various Clinical Infections: Detection of Their Drug Resistance and Virulence Determinants. Microb. Drug Resist. 2018, 24, 76–82. [Google Scholar] [CrossRef]
  81. Arthur, M.; Quintiliani, R., Jr. Regulation of VanA- and VanB-type glycopeptide resistance in enterococci. Antimicrob. Agents Chemother. 2001, 45, 375–381. [Google Scholar] [CrossRef] [Green Version]
  82. Hegstad, K.; Mikalsen, T.; Coque, T.M.; Werner, G.; Sundsfjord, A. Mobile genetic elements and their contribution to the emergence of antimicrobial resistant Enterococcus faecalis and Enterococcus faecium. Clin. Microbiol. Infect. 2010, 16, 541–554. [Google Scholar] [CrossRef]
  83. Kellogg, S.L.; Kristich, C.J. Functional Dissection of the CroRS Two-Component System Required for Resistance to Cell Wall Stressors in Enterococcus faecalis. J. Bacteriol. 2016, 198, 1326–1336. [Google Scholar] [CrossRef]
  84. Depardieu, F.; Mejean, V.; Courvalin, P. Competition between VanU(G) repressor and VanR(G) activator leads to rheostatic control of vanG vancomycin resistance operon expression. PLoS Genet. 2015, 11, e1005170. [Google Scholar] [CrossRef] [PubMed]
  85. Bloem, A.; Bax, H.I.; Yusuf, E.; Verkaik, N.J. New-Generation Antibiotics for Treatment of Gram-Positive Infections: A Review with Focus on Endocarditis and Osteomyelitis. J. Clin. Med. 2021, 10, 1743. [Google Scholar] [CrossRef]
  86. Smith, J.R.; Barber, K.E.; Raut, A.; Rybak, M.J. β-Lactams enhance daptomycin activity against vancomycin-resistant Enterococcus faecalis and Enterococcus faecium in in vitro pharmacokinetic/pharmacodynamic models. Antimicrob. Agents Chemother. 2015, 59, 2842–2848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Arias, C.A.; Panesso, D.; McGrath, D.M.; Qin, X.; Mojica, M.F.; Miller, C.; Diaz, L.; Tran, T.T.; Rincon, S.; Barbu, E.M.; et al. Genetic basis for in vivo daptomycin resistance in enterococci. N. Engl. J. Med. 2011, 365, 892–900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Stipp, R.N.; Boisvert, H.; Smith, D.J.; Höfling, J.F.; Duncan, M.J.; Mattos-Graner, R.O. CovR and VicRK regulate cell surface biogenesis genes required for biofilm formation in Streptococcus mutans. PLoS ONE 2013, 8, e58271. [Google Scholar] [CrossRef]
  89. Wu, S.; Zhang, J.; Peng, Q.; Liu, Y.; Lei, L.; Zhang, H. The Role of Staphylococcus aureus YycFG in Gene Regulation, Biofilm Organization and Drug Resistance. Antibiotics 2021, 10, 1555. [Google Scholar] [CrossRef]
  90. Voyich, J.M.; Vuong, C.; DeWald, M.; Nygaard, T.K.; Kocianova, S.; Griffith, S.; Jones, J.; Iverson, C.; Sturdevant, D.E.; Braughton, K.R.; et al. The SaeR/S gene regulatory system is essential for innate immune evasion by Staphylococcus aureus. J. Infect. Dis. 2009, 199, 1698–1706. [Google Scholar] [CrossRef] [Green Version]
  91. Karauzum, H.; Datta, S.K. Adaptive Immunity gainst Staphylococcus aureus. Curr. Top Microbiol. Immunol. 2017, 409, 419–439. [Google Scholar] [CrossRef]
  92. Qiu, Y.; Xu, D.; Xia, X.; Zhang, K.; Aadil, R.M.; Batool, Z.; Wang, J. Five major two components systems of Staphylococcus aureus for adaptation in diverse hostile environment. Microb. Pathog. 2021, 159, 105119. [Google Scholar] [CrossRef]
  93. Wu, S.; Liu, Y.; Lei, L.; Zhang, H. Endogenous antisense walR RNA modulates biofilm organization and pathogenicity of Enterococcus faecalis. Exp. Ther. Med. 2021, 21, 69. [Google Scholar] [CrossRef] [PubMed]
  94. Fakhruzzaman, M.; Inukai, Y.; Yanagida, Y.; Kino, H.; Igarashi, M.; Eguchi, Y.; Utsumi, R. Study on in vivo effects of bacterial histidine kinase inhibitor, Waldiomycin, in Bacillus subtilis and Staphylococcus aureus. J. Gen. Appl. Microbiol. 2015, 61, 177–184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Gotoh, Y.; Doi, A.; Furuta, E.; Dubrac, S.; Ishizaki, Y.; Okada, M.; Igarashi, M.; Misawa, N.; Yoshikawa, H.; Okajima, T.; et al. Novel antibacterial compounds specifically targeting the essential WalR response regulator. J. Antibiot. 2010, 63, 127–134. [Google Scholar] [CrossRef] [Green Version]
  96. Igarashi, M.; Watanabe, T.; Hashida, T.; Umekita, M.; Hatano, M.; Yanagida, Y.; Kino, H.; Kimura, T.; Kinoshita, N.; Inoue, K.; et al. Waldiomycin, a novel WalK-histidine kinase inhibitor from Streptomyces sp. MK844-mF10. J. Antibiot. 2013, 66, 459–464. [Google Scholar] [CrossRef] [Green Version]
  97. Watanabe, T.; Igarashi, M.; Okajima, T.; Ishii, E.; Kino, H.; Hatano, M.; Sawa, R.; Umekita, M.; Kimura, T.; Okamoto, S.; et al. Isolation and characterization of signermycin B, an antibiotic that targets the dimerization domain of histidine kinase WalK. Antimicrob. Agents. Chemother. 2012, 56, 3657–3663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Estrela, C.; Estrela, C.R.; Hollanda, A.C.; Decurcio Dde, A.; Pécora, J.D. Influence of iodoform on antimicrobial potential of calcium hydroxide. J. Appl. Oral Sci. 2006, 14, 33–37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Motiwala, M.A.; Habib, S.; Ghafoor, R.; Irfan, S. Comparison of antimicrobial efficacy of Calcipex and Metapex in endodontic treatment of chronic apical periodontitis: A randomised controlled trial study protocol. BMJ Open 2021, 11, e048947. [Google Scholar] [CrossRef]
  100. Babalska, Z.Ł.; Korbecka-Paczkowska, M.; Karpiński, T.M. Wound Antiseptics and European Guidelines for Antiseptic Application in Wound Treatment. Pharmaceuticals 2021, 14, 1253. [Google Scholar] [CrossRef]
  101. Okada, A.; Igarashi, M.; Okajima, T.; Kinoshita, N.; Umekita, M.; Sawa, R.; Inoue, K.; Watanabe, T.; Doi, A.; Martin, A.; et al. Walkmycin B targets WalK (YycG), a histidine kinase essential for bacterial cell growth. J. Antibiot. 2010, 63, 89–94. [Google Scholar] [CrossRef]
  102. Kato, A.; Ueda, S.; Oshima, T.; Inukai, Y.; Okajima, T.; Igarashi, M.; Eguchi, Y.; Utsumi, R. Characterization of H-box region mutants of WalK inert to the action of waldiomycin in Bacillus subtilis. J. Gen. Appl. Microbiol. 2017, 63, 212–221. [Google Scholar] [CrossRef] [Green Version]
  103. Kitayama, T.; Iwabuchi, R.; Minagawa, S.; Sawada, S.; Okumura, R.; Hoshino, K.; Cappiello, J.; Utsumi, R. Synthesis of a novel inhibitor against MRSA and VRE: Preparation from zerumbone ring opening material showing histidine-kinase inhibition. Bioorg. Med. Chem. Lett. 2007, 17, 1098–1101. [Google Scholar] [CrossRef] [PubMed]
  104. Watanabe, T.; Hashimoto, Y.; Yamamoto, K.; Hirao, K.; Ishihama, A.; Hino, M.; Utsumi, R. Isolation and characterization of inhibitors of the essential histidine kinase, YycG in Bacillus subtilis and Staphylococcus aureus. J. Antibiot. 2003, 56, 1045–1052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Lin, Y.C.; Wu, C.Y.; Huang, H.T.; Lu, M.K.; Hu, W.S.; Lee, K.T. Bacillus subtilis natto Derivatives Inhibit Enterococcal Biofilm Formation via Restructuring of the Cell Envelope. Front Microbiol. 2021, 12, 785351. [Google Scholar] [CrossRef] [PubMed]
  106. Qin, Z.; Zhang, J.; Xu, B.; Chen, L.; Wu, Y.; Yang, X.; Shen, X.; Molin, S.; Danchin, A.; Jiang, H.; et al. Structure-based discovery of inhibitors of the YycG histidine kinase: New chemical leads to combat Staphylococcus epidermidis infections. BMC Microbiol. 2006, 6, 96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Liu, H.; Zhao, Y.; Zhao, D.; Gong, T.; Wu, Y.; Han, H.; Xu, T.; Peschel, A.; Han, S.; Qu, D. Antibacterial and anti-biofilm activities of thiazolidione derivatives against clinical staphylococcus strains. Emerg. Microbes Infect. 2015, 4, e1. [Google Scholar] [CrossRef]
  108. Chen, Z.; Xiong, Y.; Tang, Y.; Zhao, Y.; Chen, J.; Zheng, J.; Wu, Y.; Deng, Q.; Qu, D.; Yu, Z. In vitro activities of thiazolidione derivatives combined with daptomycin against clinical Enterococcus faecium strains. BMC Microbiol. 2022, 22, 16. [Google Scholar] [CrossRef]
  109. Deng, X.; Zhang, C.; Chen, J.; Shi, Y.; Ma, X.; Wang, Y.; Wang, Z.; Yu, Z.; Zheng, J.; Chen, Z. Antibacterial and anti-biofilm activities of histidine kinase YycG inhibitors against Streptococcus agalactiae. J. Antibiot. 2021, 74, 874–883. [Google Scholar] [CrossRef]
  110. Furuta, E.; Yamamoto, K.; Tatebe, D.; Watabe, K.; Kitayama, T.; Utsumi, R. Targeting protein homodimerization: A novel drug discovery system. FEBS Lett. 2005, 579, 2065–2070. [Google Scholar] [CrossRef] [Green Version]
  111. Dhuri, K.; Bechtold, C.; Quijano, E.; Pham, H.; Gupta, A.; Vikram, A.; Bahal, R. Antisense Oligonucleotides: An Emerging Area in Drug Discovery and Development. J. Clin. Med. 2020, 9, 2004. [Google Scholar] [CrossRef]
  112. Wu, S.; Liu, Y.; Lei, L.; Zhang, H. Nanographene oxides carrying antisense walR RNA regulates the Enterococcus faecalis biofilm formation and its susceptibility to chlorhexidine. Lett. Appl. Microbiol. 2020, 71, 451–458. [Google Scholar] [CrossRef]
  113. Wu, S.; Liu, Y.; Zhang, H.; Lei, L. Nano-graphene oxide with antisense walR RNA inhibits the pathogenicity of Enterococcus faecalis in periapical periodontitis. J. Dent. Sci. 2020, 15, 65–74. [Google Scholar] [CrossRef] [PubMed]
  114. Moryl, M.; Palatyńska-Ulatowska, A.; Maszewska, A.; Grzejdziak, I.; Dias de Oliveira, S.; Pradebon, M.C.; Steier, L.; Różalski, A.; Poli de Figueiredo, J.A. Benefits and Challenges of the Use of Two Novel vB_Efa29212_2e and vB_Efa29212_3e Bacteriophages in Biocontrol of the Root Canal Enterococcus faecalis Infections. J. Clin. Med. 2022, 11, 6494. [Google Scholar] [CrossRef] [PubMed]
  115. Kortright, K.E.; Chan, B.K.; Koff, J.L.; Turner, P.E. Phage Therapy: A Renewed Approach to Combat Antibiotic-Resistant Bacteria. Cell Host Microbe 2019, 25, 219–232. [Google Scholar] [CrossRef] [Green Version]
  116. Merril, C.R.; Biswas, B.; Carlton, R.; Jensen, N.C.; Creed, G.J.; Zullo, S.; Adhya, S. Long-circulating bacteriophage as antibacterial agents. Proc. Natl. Acad. Sci. USA 1996, 93, 3188–3192. [Google Scholar] [CrossRef] [Green Version]
  117. Kim, K.P.; Cha, J.D.; Jang, E.H.; Klumpp, J.; Hagens, S.; Hardt, W.D.; Lee, K.Y.; Loessner, M.J. PEGylation of bacteriophages increases blood circulation time and reduces T-helper type 1 immune response. Microb. Biotechnol. 2008, 1, 247–257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Lu, T.K.; Collins, J.J. Dispersing biofilms with engineered enzymatic bacteriophage. Proc. Natl. Acad. Sci. USA 2007, 104, 11197–11202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Beaber, J.W.; Hochhut, B.; Waldor, M.K. SOS response promotes horizontal dissemination of antibiotic resistance genes. Nature 2004, 427, 72–74. [Google Scholar] [CrossRef]
  120. Chatterjee, A.; Johnson, C.N.; Luong, P.; Hullahalli, K.; McBride, S.W.; Schubert, A.M.; Palmer, K.L.; Carlson, P.E., Jr.; Duerkop, B.A. Bacteriophage Resistance Alters Antibiotic-Mediated Intestinal Expansion of Enterococci. infect. Immun. 2019, 87, e00085-19. [Google Scholar] [CrossRef] [Green Version]
  121. Duerkop, B.A.; Huo, W.; Bhardwaj, P.; Palmer, K.L.; Hooper, L.V. Molecular Basis for Lytic Bacteriophage Resistance in Enterococci. mBio 2016, 7, e01304-16. [Google Scholar] [CrossRef]
Figure 1. Structure of Zerumbone Ring Opening Derivative 2.
Figure 1. Structure of Zerumbone Ring Opening Derivative 2.
Jcm 12 00767 g001
Figure 2. Structure of Thiazolidinone Analogs Compound 2.
Figure 2. Structure of Thiazolidinone Analogs Compound 2.
Jcm 12 00767 g002
Table 1. Chemical composition of Thiazolidinone Analogs Compound 2 Derivatives.
Table 1. Chemical composition of Thiazolidinone Analogs Compound 2 Derivatives.
Thiazolidinone Analogues Compound 2 Derivatives
NameChemical Formula
H2-383-{5-{{3-(4-chlorophenyl)-2-[(4-chlorophenyl)imino]-4-oxothiazolidin-5-ylidene}methyl}furan-2-yl}benzoic acid
H2-394-{5-{{3-(4-chlorophenyl)-2-[(4-chlorophenyl)imino]-4-oxothiazolidin-5-ylidene}methyl}furan-2-yl}benzoic acid
H2-742-{4-{{3-(4-chlorophenyl)-2-[(4-chlorophenyl)imino]-4-oxothiazolidin-5-ylidene}methyl}phenoxy}acetic acid
H2-814-{5-{{3-(4-fluorophenyl)-2-[(4-phenyl)imino]-4-oxothiazolidin-5-ylidene}methyl}thiophene-2-yl} benzoic acid
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, J.; Fang, R.; Peng, Q.; Wu, S.; Lei, L. The Regulations of Essential WalRK Two-Component System on Enterococcus faecalis. J. Clin. Med. 2023, 12, 767. https://doi.org/10.3390/jcm12030767

AMA Style

Zhang J, Fang R, Peng Q, Wu S, Lei L. The Regulations of Essential WalRK Two-Component System on Enterococcus faecalis. Journal of Clinical Medicine. 2023; 12(3):767. https://doi.org/10.3390/jcm12030767

Chicago/Turabian Style

Zhang, Junqi, Rong Fang, Qi Peng, Shizhou Wu, and Lei Lei. 2023. "The Regulations of Essential WalRK Two-Component System on Enterococcus faecalis" Journal of Clinical Medicine 12, no. 3: 767. https://doi.org/10.3390/jcm12030767

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