Next Article in Journal
Synthetic Peptides against Plant Pathogenic Bacteria
Previous Article in Journal
A Method for Isolation Bacteriophage Particles-Free Genomic DNA, Exemplified by TP-84, Infecting Thermophilic Geobacillus
Previous Article in Special Issue
Targeting the Holy Triangle of Quorum Sensing, Biofilm Formation, and Antibiotic Resistance in Pathogenic Bacteria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Quorum Sensing and Quorum Quenching with a Focus on Cariogenic and Periodontopathic Oral Biofilms

by
Patricia P. Wright
1,*,† and
Srinivas Sulugodu Ramachandra
2,†
1
School of Dentistry, The University of Queensland, 288 Herston Road, Herston, QLD 4006, Australia
2
Dental Health Clinic, BC Road, Dakshina Kannada, Bantval City 574219, India
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Microorganisms 2022, 10(9), 1783; https://doi.org/10.3390/microorganisms10091783
Submission received: 6 August 2022 / Revised: 31 August 2022 / Accepted: 1 September 2022 / Published: 3 September 2022

Abstract

:
Numerous in vitro studies highlight the role of quorum sensing in the pathogenicity and virulence of biofilms. This narrative review discusses general principles in quorum sensing, including Gram-positive and Gram-negative models and the influence of flow, before focusing on quorum sensing and quorum quenching in cariogenic and periodontopathic biofilms. In cariology, quorum sensing centres on the role of Streptococcus mutans, and to a lesser extent Candida albicans, while Fusobacterium nucleatum and the red complex pathogens form the basis of the majority of the quorum sensing research on periodontopathic biofilms. Recent research highlights developments in quorum quenching, also known as quorum sensing inhibition, as a potential antimicrobial tool to attenuate the pathogenicity of oral biofilms by the inhibition of bacterial signalling networks. Quorum quenchers may be synthetic or derived from plant or bacterial products, or human saliva. Furthermore, biofilm inhibition by coating quorum sensing inhibitors on dental implant surfaces provides another potential application of quorum quenching technologies in dentistry. While the body of predominantly in vitro research presented here is steadily growing, the clinical value of quorum sensing inhibitors against in vivo oral polymicrobial biofilms needs to be ascertained.

1. Introduction

Quorum sensing allows bacteria in biofilms to chemically communicate in order to respond to environmental changes by coordinating their activity as if they were multicellular organisms [1]. Bacteria synthesize and export signalling molecules called autoinducers (AIs) [2,3]. At a particular threshold or “quorum” concentration of extracellular AI molecules, biofilm bacteria perceive the existence of these signalling compounds, resulting in changes in gene expression, and causing behavioural changes in biofilms. A cellular cascade is activated simultaneously in many bacteria which benefits the bacterial population as a whole by permitting the en masse expression of various virulence factors [4]. The property of behavioural change in biofilms due to altered gene expression because of the accumulation of AIs is known as quorum sensing (QS).
The phenomenon of QS was discovered over 50 years ago with the observation that bioluminescence in the marine bacteria Photobacterium fischeri occurred only at an increased cell density [5]. Since then, QS has been shown to be involved in the regulation of several biofilm processes including, bioluminescence, virulence, competence, conjugation, motility, and biofilm formation [6]. Various stages of biofilm development, including microcolony formation, during which extracellular polymeric substance (EPS) is secreted [7], and biofilm maturation and dispersal, are regulated by QS [8,9]. The initiation of QS can also result in the elimination of non-competent bacteria by competent bacteria, resulting in a highly virulent biofilm community [10].
Biofilm bacteria are enmeshed in a self-produced matrix of EPS, causing a many-fold increase in the resistance to the action of antimicrobials and to host immune defences compared to planktonic bacteria [11]. This feature of biofilms is becoming increasingly problematic because many chronic infections in humans are biofilm-related and are treated with antimicrobials [12]. Of particular concern is the indiscriminate overuse of antimicrobials which may result in the development of antimicrobial resistance (AMR) [13]. The rise of AMR is recognized as a major challenge currently faced by health care [13,14]. The problem is compounded by several factors including over-prescription, self-prescription and in certain countries, the unregulated prescription of antibiotics. Furthermore, the guidelines for antimicrobial stewardship are either non-existent or inconsistent [15], and pharmaceutical companies are reluctant to invest in novel antimicrobial research [16]. Additionally, a huge percentage of antimicrobials are prescribed for livestock growth promotion and in aquaculture [17,18]. Thus, researchers and clinicians have been forced to explore new ways to treat biofilm-mediated diseases. One such approach is quorum quenching (QQ) [19].
Oral diseases, including caries and periodontitis, are due to microbial dysbiosis [20]. Establishing health in such instances depends on rebiosis or establishing eubiosis [21]. However, a dysbiotic microbial flora is extremely resistant to change as it exists as a climax community. Efforts to establish eubiosis by mechanical debridement may result in short-term gains with a high chance of disease recurrence as microbial dysbiosis re-establishes [22]. To avoid disease recurrence, and to enhance the longevity of gains obtained by mechanical therapy, in some instances mechanical therapy is combined with systemic antimicrobials [22]. However, with the ever-increasing risk of the development of antimicrobial resistance, alternative therapeutic approaches need to be sought [16]. One such approach is to employ a variety of natural products such as phytochemicals [23]. The mechanism of action of specific natural products may not always be clear, but in some instances, they are known to be effective due to QS inhibition [24,25,26]. Indeed, further developing a knowledge of QS inhibition or QQ provides the opportunity to offer treatment modalities from a variety of sources including plants, bacteria, and synthetic products, free from issues of antimicrobial resistance.
QQ, which aims to interrupt bacterial communication networks responsible for virulence, has several possible applications in dentistry. Untreated dental caries in the permanent dentition accounts for 17% of global production loss [27] and represents the most prevalent condition identified in an analysis of the worldwide burden of disease [28]. Similarly, the global burden of periodontitis is enormous, and the disease has been associated with several medical conditions [29,30]. It is of little surprise then, that in recent years, research has sought alternative antimicrobial approaches [31,32], and in particular QQ has featured highly as a possible novel treatment for dental disease [26,33,34]. This narrative review aims to explain basic concepts in QS, including Gram-negative and Gram-positive intraspecies bacterial models, interspecies QS, QS in the fungus Candida albicans, and the influence of flow on QS. A further aim is to summarize the current literature regarding QS and QQ in cariogenic and periodontopathic biofilms, and to explore specific applications of QQ in implant dentistry.

2. Bacterial QS Model Systems

There are three basic systems of QS in bacteria. At an intraspecies level, Gram-negative bacteria communicate via AIs called acyl-homoserine lactones (AHLs), while in Gram-positive organisms AIs in the form of oligopeptides mediate cell-to-cell signalling. The third system for interspecies signalling, common to both Gram-positive and Gram-negative organisms, is based on a derivative of 4,5-dihydroxy-2,3-pentanedione (DHPD) called Autoinducer-2 (AI-2) [35].

2.1. Intraspecies QS in Gram-Negative Bacteria

The Gram-negative model is based on the lux regulon. The word “lux” indicates the luminescence gene [36]. In 1981, an AHL was identified as the AI for Photobacterium fischeri [37], followed by the identification in another marine species, Vibrio fischeri, of LuxR and LuxI as QS mediators. LuxR is a protein regulator and LuxI is an enzyme essential to the biosynthesis of AHL [38,39]. AHL, when in a complex with LuxR, is a transcription factor which binds to the lux regulon at a series of nucleotides known as the lux box. Binding initiates the expression of LuxI, LuxR, and proteins involved not only in bioluminescence in certain bacteria but also in bacterial virulence in particular pathogens [40]. Many Gram-negative bacteria from the phylum proteobacteria, including the well-studied pathogen Pseudomonas aeruginosa, conform to this model, each employing LuxI/LuxR homologues and their own specific AHL for QS [41]. The model is described in Figure 1.

2.2. Intraspecies QS in Gram-Positive Bacteria

The Gram-positive model resembles the Gram-negative model to some degree, in that in pathogens, the products of gene transcription result in increased bacterial virulence and produce the components needed to maintain a positive feedback loop. However, the components are specific to Gram-positive bacteria [3]. They include an autoinducing peptide (AIP) precursor, a transmembrane transport protein channel to export AIP into the extracellular environment, and a sensor histidine kinase that initiates signal transduction by causing the ATP mediated phosphorylation of a response regulator protein which then acts as a transcription factor [3]. The model is described in Figure 2.

2.3. Interspecies AI-2 Signalling Model

AI-2 is widely used by Gram-positive and -negative bacteria for interspecies communication [42]. AI-2 synthesis depends on an enzyme called LuxS which catalyses the conversion of S-ribosylhomocysteine to DHPD and homocysteine in the final step of an important bacterial pathway responsible for the methylation of proteins and nucleic acids [43]. AI-2 is a molecule that binds ribose-like structures [40]. This point is of particular relevance, as seen in Section 5.2, where sugars including D-ribose act as QS inhibitors [44]. A structural study revealed that the AI-2 in Vibrio. harveyi is a bicyclic borate derivative of DHPD, a molecule capable of adopting different cyclic configurations, depending on its environment [45]. In other bacteria, for example in the species Salmonella typhimurium, AI-2 is a monocyclic form of DHPD which does not contain boron [46].
As AI-2 accumulates extracellularly, a periplasmic receptor binds to AI-2 to initiate signal transduction that in turn enables gene transcription and virulence. In V. harveyi, this receptor protein is LuxP [45]. In several proteobacteria, the receptor protein is LsrB, which is part of an ABC transporter designed to enable AI-2 entry into the cell [40]. LsrB is transcribed from the lsr operon in response to AI-2 [40]. Recently, a third receptor was identified in Pseudomonas aeruginosa where the receptor is comprised of chemotactic and calcium channel components [47].

3. The Effect of Fluid Flow in QS

Because the process of QS occurs through AI signalling molecules which diffuse into the surrounding environment, the concentration of these signalling molecules is dependent on fluid flow [48]. Static biofilm models are still commonly encountered in in vitro research [49], although, in nature, most biofilms exist with fluid flowing in their surroundings. Thus, one of the focuses of QS research in biofilms is fluid flow. Experimental and modelling studies have demonstrated that, in a Pseudomonas aeruginosa biofilm, as the flow rate increased, higher biofilm biomass was needed to induce QS [50]. Interestingly, in some instances, an increased flow rate improves QS, whereas in other circumstances QS decreases [50]. The relationship of the flowing liquid to the biofilm is important, with QS repression in the bacteria at the biofilm–fluid interface, and QS activation in the deep biofilm layers [51]. As a similar concept, it has been demonstrated that, by a positive feedback mechanism in QS networks, bacterial biomass overcomes the gradients created by fluid flow [48]. This allows biofilms to exhibit QS-like behaviour despite fluid flow [48].

4. QS in Cariology

Dental carious lesions are associated with salivary factors such as lowered pH, causing tooth demineralization, and decreased salivary calcium and phosphate ion concentrations which are insufficient to remineralize the tooth [52]. Acidic saliva is largely linked to the presence of the acidogenic bacterial species Streptococcus mutans [53]. However, other oral pathogens including Streptococcus sobrinus and Streptococcus parasanguinis, bacteria of the genera Veillonella [53], Actinomyces and Prevotella [54], and the fungus Candida albicans, play contributory roles in carious biofilm formation [55]. In part, the dominance of S. mutans in the aetiology of dental caries helps explain the extensive QS research that exists for this species. The goal of much of this work is to identify QS inhibitors capable of attenuating dental caries [26,56,57]. From another perspective, interest is centred on S. mutans because of its role as a model organism for studying crosstalk between QS networks [58].
QS in S. mutans controls biofilm formation [34,59], genetic competence and the production of bacteriocins called mutacins [60,61,62]. Genetic competence permits S. mutans to transform its genome by the incorporation of other microorganisms’ DNA, while bacteriocins target competing bacteria to kill or inhibit them [63]. The two well-studied and interconnected QS intraspecies pathways for communication between S. mutans cells are the competence stimulating peptide (CSP) system, also known as ComABCDE, and the XIP-ComRS signalling system [10,58,63]. Recently, a further QS pathway, the PdrA/WGK system, was identified in S. mutans [64]. These three signalling systems, described in greater detail in Section 4.1.1, Section 4.1.2 and Section 4.1.3 are summarized in Figure 3.
Interspecies communication via the AI-2-LuxS system is also well documented for S. mutans biofilms [26,65,66]. Additionally, S. mutans biofilms can be attenuated by QS involving several species of commensal streptococci [67]. While the last decade has seen considerable advances in the knowledge of QS in S. mutans, it is thought that other signalling peptide systems are yet to be discovered [58].
The yeast C. albicans also plays a contributory role in dental caries. While this eukaryote is well-known for its role in the mucosal lesions of the oral candidiasis [68], its presence is positively associated with early childhood caries (EEC) [55]. Candida species are considered to be important in the maintenance of dentinal carious lesions [69], and are twice as likely to be found in the saliva of children with severe ECC compared to caries-free children [70]. In 2001, C. albicans was found to produce a QS molecule called farnesol that prevents the yeast to hyphae transition [71]. Since then, QS and QQ mechanisms have been explored in C. albicans centred on farnesol, amongst others. These are discussed in detail in Section 4.2.

4.1. QS and QQ in S. mutans

Understanding the QS networks in S. mutans provides the opportunity to utilize QQ as an alternate approach to the prevention of dental caries, by targeting signalling networks and the pathogenicity of S. mutans. Typically, experimental studies are performed in vitro and employ reverse transcription and polymerase chain reaction assays to demonstrate the downregulation of genes involved in QS. This is accompanied by phenotypic change such as a decrease in biofilm volume, hydrophobicity or EPS [26,34,59]. Some studies provide further evidence of QQ by the inclusion of assays demonstrating reductions in genetic transformation and bacteriocin production [56,59].
The CSP-ComABCDE, XIP-ComRS and the interspecies AI-2/LuxS pathways are all targets for QQ. Interference with the AI-2/LuxS pathway is usually studied using mutant strains deficient in luxS [65,66]. In the case of biofilms of ΔluxS S. gordonii coaggregates with S. mutansWT, the lack of S. gordonii AI-2 signalling results in decreases in the dual-species biofilm mass [65]. In ΔluxS S. mutans strains, ABC transporters are downregulated [66]. QQ inhibition of S. mutans biofilms has been achieved using substances derived from sources as diverse as natural products, bacterially produced antimicrobials, synthetic materials, and human saliva (Figure 3).

4.1.1. CSP/ComABCDE QS Pathway in S. mutans

Over 20 years ago, a competence-stimulating peptide export system based on an ATP binding cassette (ABC) transporter, ComAB, was described for S. mutans [72]. Soon after, a QS feedback loop, ComCDE was reported [73]. The system follows the Gram-positive model outlined in Section 2.2. The comC, comD and comE genes express, respectively, ComC, the precursor of the competence stimulating peptide (CSP), ComD, a transmembrane histidine kinase signal transducer and ComE, a response regulator protein [73]. ComC is first truncated by the ABC transporter into a 21 amino acid protein, then further processed by an extracellular membrane-bound protease, SepM, into CSP, an 18 amino acid AI [74]. CSP binding of ComD results in ComE phosphorylation which causes the transcription of an alternate sigma factor, called ComX, from the comX gene to induce an S. mutans phenotype that is genetically competent [73]. The ComABCDE system is also important for the initiation of biofilm formation [75], and the expression of bacteriocins from, for example, the bacteriocin expressing genes nlmA and nlmC [76].
Additionally, it is known that the ComABCDE system can be bypassed because both competence and bacteriocin production can be activated independently through a response regulator called HdrR [60]. HdrR is expressed from the hdrRM operon which also expresses HdrM, an inhibitor protein that blocks the action of HdrR, thus providing alternative regulation of the conversion to genetically competent and bacteriocin-producing phenotypic forms [60].

4.1.2. XIP-ComRS QS Pathway in S. mutans

The ComABCDE system is networked with the other main feedback loop for intercellular communication between cells in S. mutans biofilms which involves the comX inducing peptide (XIP), expressed as the percussor ComS [77]. A significant difference between the two pathways, is that in addition to transmembrane transporters, cell lysis appears to be a major mechanism for discharging the AI extracellularly [77]. XIP is then imported back into the cell by a membrane oligopeptide permease, OppD [78], where, in many streptococcal species including S. mutans, it binds the response regulator ComR [79]. The ComR-XIP complex acts as a transcriptional activator of the genes comS, comR and comX in a feedback loop, with transcription of comX resulting in genetic competence [61]. Current work in other streptococci has identified key amino acids in the interaction between ComR and XIP which results in the formation of an active complex [80].

4.1.3. PdrA-WGK QS Pathway in S. mutans

In 2021, the S. mutans wgk operon was shown to be regulated by a short hydrophobic AIP (SHP) and a response regulator called PdrA. SHP is exported out of the cell by the transmembrane exporter PptAB and reimported by an Opp transporter. Activation of the operon results in the production of a peptide called tryglysin B, a 7-amino acid peptide containing the sequence Try-Gly-Lys. Tryglysin B inhibits several oral streptococci including S. mitis, S. oralis, S. sanguinis, and to a lesser extent S. gordonii. Tryglsin B is also autoinhibitory; however, S. mutans recovers from this inhibition by 24 h [64]. Detailing the PdrA-WGK pathway has further helped understand virulence in S. mutans by showing how it can inhibit competing oral streptococci.
Figure 3. The CSP/ComABCDE, XIP-ComRS and PdrA-WGK pathways and their interaction, inhibition, and activation in S. mutans.
Figure 3. The CSP/ComABCDE, XIP-ComRS and PdrA-WGK pathways and their interaction, inhibition, and activation in S. mutans.
Microorganisms 10 01783 g003

4.1.4. QQ in S. mutans by Plant Products

Inhibition of S. mutans biofilm formation, in association with downregulation of comDE, has been associated with the treatment of biofilms with several plant products. These include baicalein (5,6,7-trihydroxyflavone), obtained from Scutellaria baicalensis and Scutellaria lateriflora [26], extracts from Rhodiola rosea [81], and Emblica officinalis, the Indian gooseberry [24]. In the case of baicalein and E. officinalis, biofilm hydrophobicity, important to bacteria-carbohydrate interactions, was shown to decrease. Baicalein treatment also resulted in the downregulation of comX and luxS, illustrating the potential to disrupt both intraspecies and interspecies cellular communication [26]. Combining plant products with common antimicrobials and fluoride has been tested. When chlorhexidine and fluoride were coupled with trans-cinnamaldehyde obtained from cinnamon, downregulation of comADE, comX and luxS was observed, associated with biofilm inhibition which complemented the action of chlorhexidine [82].

4.1.5. QQ in S. mutans by Microbial Products

S. mutans competitively coexists in oral biofilms with several other species, the products of which can disrupt QS in S. mutans. S. gordonii and S. sanguinis can not only be inhibited by S. mutans produced tryglysin B, but conversely, can disrupt ComRS and ComE signalling in S. mutans, causing the degradation of S. mutans biofilms [67]. Walkmycin C, an antibiotic produced by Streptomyces strain MK632-100F11 [83], targets S. mutans comDE [84]. Fructanase, an enzyme produced by Streptococcus salivarius, [85] adversely affects S. mutans competence and bacteriocin production [59]. The products of probiotic Lactobacillus species can downregulate S. mutans comD [86], while antimicrobials produced by Lactiplantibacillus species target S. mutans, comA and comX [57]. Thus, there appear to be many opportunities for bacterial products to be trialled as anticaries agents. However, it must also be recognized that a complex competitive bacterial network is operational in the oral cavity, the implications of which are far from completely understood.

4.1.6. QQ in S. mutans by Synthetic Molecules

Synthetic QQ molecules are a further avenue of research in caries prevention. A bicyclic amine of the quinuclidine class was found to downregulate the bacteriocin-producing genes nlmA and nlmC and inhibit genetic competence in S. mutans. Inhibition occurred via non-competitive binding of the peptidase domain of ComA, causing allosteric changes at the enzyme active site [76]. Other synthetic compounds include small peptide analogues of XIP, capable of disrupting XIP binding to ComR [56], and nano-quercetin particles used in conjunction with photodynamic therapy, causing downregulation of comBCDE [87]. Finally, human saliva contains a naturally occurring mucin, MUC5B, able to inhibit S. mutans biofilm formation. The mechanism operates through the linkages and structures of the O-linked glycans forming the mucin. It affects both the CSP and XIP pathways [34].

4.2. QS and QQ in Candida albicans and Dual Species S. mutans-C. albicans Biofilms

Farnesol is a lipophilic, QS molecule that is secreted into the extracellular environment by C. albicans [71]. At high cell density, farnesol suppresses an enzymatic pathway, the Ras1-Cdc35-PKA-Efg1 pathway, that permits hyphal transformation important to the virulence of this species [88]. Farnesol has a sesquiterpene structure similar to that of dodecanal which also is able to repress the same pathway and inhibit the transformation to the hyphal form [88]. Another compound, the tryrosine derivative tyrosol, like farnesol, is a QS molecule in C. albicans, and also like farnesol, it can inhibit C. albicans biofilm formation [89].
However, at low concentrations, farnesol can promote rather than inhibit biofilm formation important in cariology by stimulating S. mutans growth [90]. In this context, farnesol secreted by C. albicans can augment biofilm growth in dual-species biofilms of C. albicans and S. mutans by increasing the expression of S. mutans glucosyl transferases [90]. Glucosyl transferases increase sugar metabolism in S. mutans, resulting in the secretion of a glucan-rich adherent plaque [91], favourable to C. albicans [90]. This interkingdom relationship between S. mutans and C. albicans appears to be finely balanced, but symbiotic. It permits C. albicans to grow well, when unaccompanied it would not, and at the same time, it increases the virulence of S. mutans [90]. It has been shown that this relationship induces the transcription of S. mutans comC and comX, implicating the AIs CSP and XIP, genetic competence and bacteriocins in the mechanism [92]. The involvement of S. mutans QS networks in these dual-species biofilms was confirmed by a later study which also reported changes to comC expression [93].
It is noteworthy, however, that at high farnesol concentrations in S. mutans-C. albicans. biofilms, growth is inhibited [90], providing opportunities to employ QQ against cariogenic biofilms. Indeed, farnesol has been used in combination with fluoride and plant products (a flavonoid called myricetin and flavonoid metabolites) to reduce S. mutans-C. albicans cell numbers [94]. Other plant products have been used to inhibit dual species S. mutans-C. albicans biofilms. Both curcumin [25] and the plant derivative thymol [95] interfere with S. mutans CSP QS via inhibition of comCD expression. This impairs biofilm development, an effect which, in the case of curcumin, is stronger for S. mutans than for C. albicans. Certain drugs have also been used as QS inhibitors in this situation. Thiazolidinediones are used in the treatment of diabetes [96]. A related compound, thiazolidinedione-8 has been shown to be effective against QS in S. mutans-C. albicans biofilms, with the effect being strongest against the C. albicans component [97].

5. QS and QQ in Periodontitis

The oral cavity is known to harbour more than 700 to 800 diverse bacterial species [57]. A complex interplay exists among this large group of bacteria during health and disease. The microbial interaction is not merely limited to an interaction between healthy and pathogenic bacteria, but also involves a positive feedback loop between the bacteria and the host immune-inflammatory response [20]. Thus, during eubiosis and dysbiosis, the interaction is multi-dimensional [20]. Some of the bacterial by-products generated by one group of bacteria may be sensed as growth promoters by other groups of bacteria, while, for another group of bacteria, the same by-products may be inhibitory. Several periodontal pathogens produce AI-2 molecules which may be involved in cell–cell signalling within these biofilms [20]. AI-2 signalling influences the response to host-generated stress, environmental stress, the acquisition of iron and hemin and the development of periodontal biofilms [20]. Sites with progressive periodontitis can demonstrate increased gene expression related to periodontal pathogenesis. This particularly applies to the oxidative stress response, ferrous iron transport, amino-acid transport, and the synthesis of lipopolysaccharides [98]. Non-progressive sites may exhibit stable gene expression over the follow-up [98].
Several periodontal bacteria including Porphyromonas gingivalis, Treponema denticola, Tannerella forsythia, Prevotella intermedia and Fusobacterium nucleatum have been studied for their QS properties and may serve as potential targets for QQ [99]. Because biofilm formation and various virulence properties of periodontal biofilms are controlled by QS, several possibilities exist to interfere with QS signalling systems, and numerous novel attempts at QQ have been made.
The majority of the research investigating QS among bacteria associated with periodontitis is focused on P. gingivalis, a putative bacterium which is considered a “keystone pathogen” in periodontitis [100]. Its presence in sufficient quantities is known to convert the entire microbiota into dysbiosis [20,101]. The red-complex species, P. gingivalis, T. denticola and T. forsythia, which are frequently isolated from sites with active periodontitis [102], form an important group of pathogens evaluated for their QS behaviour. F. nucleatum, often considered as the bridge organism responsible for the coaggregation between early commensal bacteria and late pathogenic bacteria [103], has also been studied for QS properties, and subsequently as a target for QQ in the treatment of periodontitis. While much of the QS and QQ research in periodontitis is in in vitro models, a limited number of in vivo models exist, especially in murine species.

5.1. QS Interactions among Periodontal Pathogens

Quorum sensing also influences the complex intraspecies and interspecies interactions among various pathogens. In 2001, strains of P. intermedia, F. nucleatum, and P. gingivalis were reported as having AI-2-like activities [99]. The authors also reported the inability of these pathogens to produce homologues of AHL [99]. However, recent in vitro evidence suggests that periodontal pathogens produce short-chain AHL molecules [104]. In subgingival biofilms, there are reports that the AI-2 of F. nucleatum promoted biofilm formation and helped in inter-species coaggregation between F. nucleatum and each of the red complex species [105]. Similarly, F. nucleatum AI-2 resulted in increased biofilm formation of S. gordonii and S. oralis [106], and enhanced the attachment of F. nucleatum to these streptococci [105]. The strongest interaction existed between F. nucleatum and S. mitis and the addition of other bacteria may interfere with this co-adherence [107]. Aggregatibacter actinomycetemcomitans serotypes b and f were shown to aggregate with F. nucleatum, mediated by QS through an O-polysaccharide [108].
Eikenella corrodens can express a gene that is essential for the production of AI-2 molecules and the presence of AI-2 molecules can influence the formation of E. corrodens biofilms [109]. The efficiency of E. corrodens biofilm formation in the presence of AI-2 molecules was 1.3 times greater than in its absence [109]. In dual-species biofilms grown on artificial saliva, Aggregatibacter actinomycetemcomitans activated the QS regulon of S. mutans [110]. Biofilm formation and the virulence of Aggregatibacter actinomycetemcomitans are controlled by the release of AI-2. However, gene expression following the release of AI-2 molecules is controlled by the release of the QseBC two-component system as evidenced in an in vivo mouse model [111].
Several in vitro and in vivo models have been explored to understand the network of QS in periodontitis. However, due to the complex interplay of various bacteria involved in periodontitis, more research work is essential to decipher the role of QS in the pathogenesis of periodontitis.

5.2. QQ Compounds in Periodontitis

5.2.1. Furanone Compounds and Combinations with D-Ribose

Furanones, including 2(5H)-furanone, have been explored as potential QQ agents against periodontal pathogens [112]. This is because various furanones are similar to N-AHL in terms of their chemical structure. Additionally, they inhibit biofilm formation without the development of AMR [112].
Brominated furanones show promise as QQ compounds. Monocyclic brominated furanones produced by the red algae Delisea pulchra, have antibiofilm properties associated with the modulation of QS genes [113]. Synthetic bicyclic brominated furanones have applications against F. nucleatum, P. gingivalis, and T. forsythia biofilms, as evidenced by reductions in biofilm biomass and thickness, [114]. The further potential of the use of brominated furanones against periodontal pathogens is shown by their ability to reduce P. gingivalis biofilm biomass, causing the scattering of P. gingivalis cells [115]. The authors reported that brominated furanones showed concentration-dependent effects and had potential as an antimicrobial treatment for periodontitis [115].
Brominated furanones have also been combined with D-ribose to evaluate their anti-AI-2 effect on coaggregates of P. gingivalis and F. nucleatum in an in vivo murine model [44]. A reduction in the periodontal breakdown, as evidenced by reduced alveolar bone loss and decreased bacterial infection of the periodontal tissues was observed [44]. Similarly, the combination of a brominated furanone [116] and D-ribose in a murine model showed reduced bone loss and decreased levels of P. gingivalis [117]. Furthermore, (5Z)-4-bromo-5-(bromomethylene)-2(5H)-furanone and D-ribose, can inhibit AI-2 induced F. nucleatum coaggregation, with individually, all three red complex pathogens [105]. These same QS inhibitors were also shown to interfere with early oral biofilm colonization by streptococci, again via the inhibition of F. nucleatum AI-2 [106].

5.2.2. D-Galactose

D-galactose also has QQ properties. The results of crystal violet and confocal microscopy assays showed that D-galactose inhibited the activity of AI-2. This resulted in decreased formation of F. nucleatum, P. gingivalis, and T. forsythia biofilms [118]. D-galactose also inhibited the coaggregation of A. actinomycetemcomitans serotypes b and f with F. nucleatum [108].

5.2.3. Natural Products

Several plant-derived products and phenols, such as coumarin and green tea extracts, have been explored for their QS inhibition properties. Coumarin, inhibited biofilm formation at various stages and downregulated the expression of P. gingivalis biofilms and luxS genes. Extracts from green tea and its main component epigallocatechin-3-gallate (EGCG) inhibited the growth and adherence of P. gingivalis biofilms and also downregulated various genes responsible for the virulence of P. gingivalis [119]. While green tea extract and ECCG were able to inhibit AI-2 in a Vibrio harveyi model, their role in QQ in P. gingivalis biofilms is yet to be verified [119].

5.2.4. Synthetic Analogues of AHL

Several synthetic analogues of AHLs have been shown to inhibit P. gingivalis biofilm formation with three of them able to significantly decrease the number of biofilm cells [120]. A less well-formed biofilm structure, as demonstrated by confocal microscopy, was observed. Modifications to AHL were made by the introduction of varying numbers of either alcohol or amine groups into the acyl groups, which also contained altered numbers of carbon atoms.

6. Coating Surfaces of Dental Implants with QS Inhibitors

Every year, many medical devices including dental implants, prostheses, bone plates, bone grafts and membranes are placed within the mouth [121,122,123]. The abutment surfaces of dental implants typically get covered by biofilms, leading to medical-device-associated infections such as peri-implantitis [124,125]. Thus, device removal or additional surgeries are required [126]. Recently, alternative approaches including the use of lasers and photobiomodulation have reported significant adjunctive clinical benefits compared to conventional scaling and root debridement in the treatment of peri-implantitis [127]. QQ may provide a solution to this problem. Coating the dental implant abutment surfaces with QS inhibitors may result in reduced biofilm formation (Figure 4).
Microparticles coated with QS inhibitors (calcium-binding polymer-coated poly(lactide-co-glycolide) have been placed on hydroxyapatite surfaces to prevent S. mutans biofilm formation [128]. Because most dental hard tissues are made up of hydroxyapatite, these inhibitors may have the potential for research in other dental disciplines.
Pre-treatment of titanium discs with D-arabinose for three minutes has been shown to inhibit the formation of both monospecies biofilms of S. oralis, F. nucleatum and P. gingivalis and a biofilm consortium of the same three bacterial species [129]. The authors also suggested that D-arabinose had initial anti-adhesive properties against biofilm-forming bacteria and sustained inhibitory activity against AI-2 [129]. The research concluded that D-arabinose may be used to prevent peri-implant mucositis and peri-implantitis [129].
QS inhibition has been demonstrated in other fields. Agents involved in urinary tract infections, Pseudomonas aeruginosa and Serratia marcescens, can be inhibited by coating thiazolinyl-picolinamide-based palladium (II) complexes on urinary catheters [130]. Similarly, heteroleptic pincer palladium (II) complex coated orthopaedic implants resulted in QQ and inhibited the development of Acinetobacter baumannii biofilms [131]. However, the possible application of this approach to peri-implantitis may pose problems. An important differentiating factor in urinary tract and orthopaedic infections versus peri-implantitis is that the former are mostly due to single species or dual-bacterial species. Conversely, peri-implantitis is associated with a high bacterial diversity [132,133].

7. Conclusions, Limitations, and Future Directions

Improving the knowledge base of QS and QQ mediators and mechanisms has the potential for the development of antibiofilm treatments. Some areas of dentistry are underrepresented in the QS/QQ literature, and the focus of this review aligns with the main areas of research. To date, in cariology and periodontology, interfering with QS signals has been effective predominantly in in vitro research to decrease biofilm formation and virulence. The switch to in vivo focused research would be the next step in realizing some of the clinical potential for QS inhibitors from diverse natural and synthetic substances, including carbohydrates, and autoinducer analogues. Such research would reveal the effects of QQ substances on naturally occurring polymicrobial biofilms. Many of the in vitro studies presented in this review have been performed on monospecies biofilms, particularly with respect to dental caries, where S. mutans biofilms predominate. Periodontal QQ research is based on P. gingivalis or a consortium of bacteria including, amongst others, P. gingivalis, T. forsythia and F. nucleatum. However, current evidence indicates that polymicrobial synergy and dysbiosis of the entire microbial community are involved in dental diseases. Additionally, naturally occurring biofilm infections are a result of complex interactions between multiple pathogenic bacteria, host-compatible bacteria, and the host’s immuno-inflammatory cascade. Thus, whether the positive QQ results outlined in this review can be replicated in naturally occurring microcosms remains to be ascertained.

Author Contributions

All authors contributed to the content, drafting and writing of this document. 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.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grandclément, C.; Tannières, M.; Moréra, S.; Dessaux, Y.; Faure, D. Quorum quenching: Role in nature and applied developments. FEMS Microbiol. Rev. 2016, 40, 86–116. [Google Scholar] [CrossRef] [PubMed]
  2. Bzdrenga, J.; Daudé, D.; Rémy, B.; Jacquet, P.; Plener, L.; Elias, M.; Chabrière, E. Biotechnological applications of quorum quenching enzymes. Chem. Biol. Interact. 2017, 267, 104–115. [Google Scholar] [CrossRef] [PubMed]
  3. Rutherford, S.T.; Bassler, B.L. Bacterial quorum sensing: Its role in virulence and possibilities for its control. Cold Spring Harb. Perspect. Med. 2012, 2, a012427. [Google Scholar] [CrossRef] [PubMed]
  4. Whiteley, M.; Diggle, S.P.; Greenberg, E.P. Bacterial quorum sensing: The progress and promise of an emerging research area. Nature 2017, 551, 313. [Google Scholar] [CrossRef] [PubMed]
  5. Nealson, K.H.; Platt, T.; Hastings, J.W. Cellular control of the synthesis and activity of the bacterial luminescent system. J. Bacteriol. 1970, 104, 313–322. [Google Scholar] [CrossRef] [PubMed]
  6. Miller, M.B.; Bassler, B.L. Quorum sensing in bacteria. Annu. Rev. Microbiol. 2001, 55, 165–199. [Google Scholar] [CrossRef]
  7. Jayathilake, P.G.; Jana, S.; Rushton, S.; Swailes, D.; Bridgens, B.; Curtis, T.; Chen, J. Extracellular polymeric substance production and aggregated bacteria colonization influence the competition of microbes in biofilms. Front. Microbiol. 2017, 8, 1865. [Google Scholar] [CrossRef]
  8. Frederick, M.R.; Kuttler, C.; Hense, B.A.; Eberl, H.J. A mathematical model of quorum sensing regulated EPS production in biofilm communities. Theor. Biol. Med. Model. 2011, 8, 8. [Google Scholar] [CrossRef]
  9. Tan, C.H.; Koh, K.S.; Xie, C.; Tay, M.; Zhou, Y.; Williams, R.; Ng, W.J.; Rice, S.A.; Kjelleberg, S. The role of quorum sensing signalling in EPS production and the assembly of a sludge community into aerobic granules. ISME J. 2014, 8, 1186–1197. [Google Scholar] [CrossRef]
  10. Smith, E.G.; Spatafora, G.A. Gene regulation in S. mutans: Complex control in a complex environment. J. Dent. Res. 2012, 91, 133–141. [Google Scholar] [CrossRef]
  11. Balcázar, J.L.; Subirats, J.; Borrego, C.M. The role of biofilms as environmental reservoirs of antibiotic resistance. Front. Microbiol. 2015, 6, 1216. [Google Scholar] [CrossRef] [PubMed]
  12. Hall-Stoodley, L.; Costerton, J.W.; Stoodley, P. Bacterial biofilms: From the natural environment to infectious diseases. Nat. Rev. Microbiol. 2004, 2, 95–108. [Google Scholar] [CrossRef] [PubMed]
  13. Shelenkov, A. Whole-genome sequencing of pathogenic bacteria-new insights into antibiotic resistance spreading. Microorganisms 2021, 9, 2624. [Google Scholar] [CrossRef] [PubMed]
  14. Aslam, B.; Khurshid, M.; Arshad, M.I.; Muzammil, S.; Rasool, M.; Yasmeen, N.; Shah, T.; Chaudhry, T.H.; Rasool, M.H.; Shahid, A.; et al. Antibiotic resistance: One health one world outlook. Front. Cell Infect. Microbiol. 2021, 11, 771510. [Google Scholar] [CrossRef]
  15. Collaborators. Global burden of bacterial antimicrobial resistance in 2019: A systematic analysis. Lancet 2022, 399, 629–655. [Google Scholar] [CrossRef]
  16. Piddock, L.J. The crisis of no new antibiotics—What is the way forward? Lancet Infect. Dis. 2012, 12, 249–253. [Google Scholar] [CrossRef]
  17. Albernaz-Gonçalves, R.; Olmos, G.; Hötzel, M.J. Exploring farmers’ reasons for antibiotic use and misuse in pig farms in brazil. Antibiotics 2021, 10, 331. [Google Scholar] [CrossRef]
  18. Gozdzielewska, L.; King, C.; Flowers, P.; Mellor, D.; Dunlop, P.; Price, L. Scoping review of approaches for improving antimicrobial stewardship in livestock farmers and veterinarians. Prev. Vet. Med. 2020, 180, 105025. [Google Scholar] [CrossRef]
  19. Roy, V.; Adams, B.L.; Bentley, W.E. Developing next generation antimicrobials by intercepting AI-2 mediated quorum sensing. Enzym. Microb. Technol. 2011, 49, 113–123. [Google Scholar] [CrossRef]
  20. Hajishengallis, G.; Lamont, R.J. Polymicrobial communities in periodontal disease: Their quasi-organismal nature and dialogue with the host. Periodontology 2000 2021, 86, 210–230. [Google Scholar] [CrossRef]
  21. Iebba, V.; Totino, V.; Gagliardi, A.; Santangelo, F.; Cacciotti, F.; Trancassini, M.; Mancini, C.; Cicerone, C.; Corazziari, E.; Pantanella, F.; et al. Eubiosis and dysbiosis: The two sides of the microbiota. New Microbiol. 2016, 39, 1–12. [Google Scholar] [PubMed]
  22. Magda Feres, H.D. Systemic Antibiotics in Periodontal Therapy. In Lindhe’s Clinical Periodontology and Implant Dentistry; Niklaus, W.V.G., Lang, P., Mariano Sanz, T.B., Eds.; Wiley-Blackwell: Hoboken, NJ, USA, 2021. [Google Scholar]
  23. Ali, I.A.A.; Neelakantan, P. Antibiofilm activity of phytochemicals against Enterococcus faecalis: A literature review. Phytother. Res. 2022, 36, 2824–2838. [Google Scholar] [CrossRef] [PubMed]
  24. Hasan, S.; Danishuddin, M.; Adil, M.; Singh, K.; Verma, P.K.; Khan, A.U. Efficacy of E. officinalis on the cariogenic properties of Streptococcus mutans: A novel and alternative approach to suppress quorum-sensing mechanism. PLoS ONE 2012, 7, e40319. [Google Scholar]
  25. Li, X.; Yin, L.; Ramage, G.; Li, B.; Tao, Y.; Zhi, Q.; Lin, H.; Zhou, Y. Assessing the impact of curcumin on dual-species biofilms formed by Streptococcus mutans and Candida albicans. Microbiologyopen 2019, 8, e937. [Google Scholar] [CrossRef]
  26. Vijayakumar, A.; Sarveswari, H.B.; Vasudevan, S.; Shanmugam, K.; Solomon, A.P.; Neelakantan, P. Baicalein inhibits Streptococcus mutans biofilms and dental caries-related virulence phenotypes. Antibiotics 2021, 10, 215. [Google Scholar] [CrossRef]
  27. Listl, S.; Galloway, J.; Mossey, P.A.; Marcenes, W. Global economic impact of dental diseases. J. Dent. Res. 2015, 94, 1355–1361. [Google Scholar] [CrossRef]
  28. Kassebaum, N.; Bernabé, E.; Dahiya, M.; Bhandari, B.; Murray, C.; Marcenes, W. Global burden of untreated caries: A systematic review and metaregression. J. Dent. Res. 2015, 94, 650–658. [Google Scholar] [CrossRef]
  29. Das, M.; Upadhyaya, V.; Ramachandra, S.S.; Jithendra, K.D. Periodontal treatment needs in diabetic and non-diabetic individuals: A case-control study. Indian J. Dent. Res. 2011, 22, 291–294. [Google Scholar] [CrossRef]
  30. Wu, L.; Zhang, S.Q.; Zhao, L.; Ren, Z.H.; Hu, C.Y. Global, regional and national burden of periodontitis from 1990 to 2019: Results from the Global Burden of Disease study 2019. J. Periodontol. 2022. Online ahead of print. Available online: https://aap.onlinelibrary.wiley.com/doi/10.1002/JPER.21-0469 (accessed on 31 August 2022).
  31. Mendhi, J.; Ramachandra, S.S.; Prasadam, I.; Ivanovski, S.; Yang, Y.; Xiao, Y. Endogenous nitric oxide-generating surfaces via polydopamine-copper coatings for preventing biofilm dispersal and promoting microbial killing. Mater. Sci. Eng. C Mater. Biol. Appl. 2021, 128, 112297. [Google Scholar] [CrossRef]
  32. Xiang, E.; Gómez-Cerezo, M.N.; Ali, Y.; Ramachandra, S.S.; Yang, N.; Dargusch, M.; Moran, C.S.; Ivanovski, S.; Abdal-Hay, A. Surface Modification of pure zinc by acid etching: Accelerating the corrosion rate and enhancing biocompatibility and antibacterial characteristics. ACS Appl. Mater. Interfaces 2022, 14, 22554–22569. [Google Scholar] [CrossRef]
  33. Kim, W.J.; Soh, Y.; Heo, S.-M. Recent advances of therapeutic targets for the treatment of periodontal disease. Biomol. Ther. 2021, 29, 263. [Google Scholar] [CrossRef] [PubMed]
  34. Werlang, C.A.; Chen, W.G.; Aoki, K.; Wheeler, K.M.; Tymm, C.; Mileti, C.J.; Burgos, A.C.; Kim, K.; Tiemeyer, M.; Ribbeck, K. Mucin O-glycans suppress quorum-sensing pathways and genetic transformation in Streptococcus mutans. Nat. Microbiol. 2021, 6, 574–583. [Google Scholar] [CrossRef] [PubMed]
  35. Duddy, O.P.; Bassler, B.L. Quorum sensing across bacterial and viral domains. PLoS Pathog. 2021, 17, e1009074. [Google Scholar] [CrossRef] [PubMed]
  36. Fuqua, W.C.; Winans, S.C.; Greenberg, E.P. Quorum sensing in bacteria: The LuxR-LuxI family of cell density-responsive transcriptional regulators. J. Bacteriol. 1994, 176, 269–275. [Google Scholar] [CrossRef] [PubMed]
  37. Eberhard, A.; Burlingame, A.L.; Eberhard, C.; Kenyon, G.L.; Nealson, K.H.; Oppenheimer, N. Structural identification of autoinducer of Photobacterium fischeri luciferase. Biochemistry 1981, 20, 2444–2449. [Google Scholar] [CrossRef]
  38. Engebrecht, J.; Nealson, K.; Silverman, M. Bacterial bioluminescence: Isolation and genetic analysis of functions from Vibrio fischeri. Cell 1983, 32, 773–781. [Google Scholar] [CrossRef]
  39. Engebrecht, J.; Silverman, M. Identification of genes and gene products necessary for bacterial bioluminescence. Proc. Natl. Acad. Sci. USA 1984, 81, 4154–4158. [Google Scholar] [CrossRef]
  40. Papenfort, K.; Bassler, B.L. Quorum sensing signal-response systems in Gram-negative bacteria. Nat. Rev. Microbiol. 2016, 14, 576–588. [Google Scholar] [CrossRef]
  41. Case, R.J.; Labbate, M.; Kjelleberg, S. AHL-driven quorum-sensing circuits: Their frequency and function among the Proteobacteria. ISME J. 2008, 2, 345–349. [Google Scholar] [CrossRef]
  42. Bassler, B.L.; Greenberg, E.P.; Stevens, A.M. Cross-species induction of luminescence in the quorum-sensing bacterium Vibrio harveyi. J. Bacteriol. 1997, 179, 4043–4045. [Google Scholar] [CrossRef]
  43. Schauder, S.; Shokat, K.; Surette, M.G.; Bassler, B.L. The LuxS family of bacterial autoinducers: Biosynthesis of a novel quorum-sensing signal molecule. Mol. Microbiol. 2001, 41, 463–476. [Google Scholar] [CrossRef] [PubMed]
  44. Ben Amara, H.; Song, H.Y.; Ryu, E.; Park, J.S.; Schwarz, F.; Kim, B.M.; Choi, B.K.; Koo, K.T. Effects of quorum-sensing inhibition on experimental periodontitis induced by mixed infection in mice. Eur. J. Oral Sci. 2018, 126, 449–457. [Google Scholar] [CrossRef] [PubMed]
  45. Chen, X.; Schauder, S.; Potier, N.; Van Dorsselaer, A.; Pelczer, I.; Bassler, B.L.; Hughson, F.M. Structural identification of a bacterial quorum-sensing signal containing boron. Nature 2002, 415, 545–549. [Google Scholar] [CrossRef] [PubMed]
  46. Miller, S.T.; Xavier, K.B.; Campagna, S.R.; Taga, M.E.; Semmelhack, M.F.; Bassler, B.L.; Hughson, F.M. Salmonella typhimurium recognizes a chemically distinct form of the bacterial quorum-sensing signal AI-2. Mol. Cell 2004, 15, 677–687. [Google Scholar] [CrossRef]
  47. Zhang, L.; Li, S.; Liu, X.; Wang, Z.; Jiang, M.; Wang, R.; Xie, L.; Liu, Q.; Xie, X.; Shang, D.; et al. Sensing of autoinducer-2 by functionally distinct receptors in prokaryotes. Nat. Commun. 2020, 11, 5371. [Google Scholar] [CrossRef]
  48. Dalwadi, M.P. Pearce, P. Emergent robustness of bacterial quorum sensing in fluid flow. Proc. Natl. Acad. Sci. USA 2021, 118, e2022312118. [Google Scholar] [CrossRef]
  49. Azeredo, J.; Azevedo, N.F.; Briandet, R.; Cerca, N.; Coenye, T.; Costa, A.R.; Desvaux, M.; Di Bonaventura, G.; Hebraud, M.; Jaglic, Z.; et al. Critical review on biofilm methods. Crit. Rev. Microbiol. 2017, 43, 313–351. [Google Scholar] [CrossRef]
  50. Kirisits, M.J.; Margolis, J.J.; Purevdorj-Gage, B.L.; Vaughan, B.; Chopp, D.L.; Stoodley, P.; Parsek, M.R. Influence of the hydrodynamic environment on quorum sensing in Pseudomonas aeruginosa biofilms. J. Bacteriol. 2007, 189, 8357–8360. [Google Scholar] [CrossRef]
  51. Kim, M.K.; Ingremeau, F.; Zhao, A.; Bassler, B.L.; Stone, H.A. Local and global consequences of flow on bacterial quorum sensing. Nat. Microbiol. 2016, 1, 15005. [Google Scholar] [CrossRef]
  52. Nicolae, V.; Neamtu, B.; Picu, O.; Stefanache, M.A.M.; Cioranu, V.S.I. The comparative evaluation of salivary biomarkers (Calcium, Phosphate, Salivary pH) in cario-resistance versus cario-activity. Rev. Chim. 2016, 67, 821–824. [Google Scholar]
  53. Gross, E.L.; Beall, C.J.; Kutsch, S.R.; Firestone, N.D.; Leys, E.J.; Griffen, A.L. Beyond Streptococcus mutans: Dental caries onset linked to multiple species by 16S rRNA community analysis. PLoS ONE 2012, 7, e47722. [Google Scholar] [CrossRef] [PubMed]
  54. Johansson, I.; Witkowska, E.; Kaveh, B.; Lif Holgerson, P.; Tanner, A.C. The microbiome in populations with a low and high prevalence of caries. J. Dent. Res. 2016, 95, 80–86. [Google Scholar] [CrossRef] [PubMed]
  55. Xiao, J.; Huang, X.; Alkhers, N.; Alzamil, H.; Alzoubi, S.; Wu, T.T.; Castillo, D.A.; Campbell, F.; Davis, J.; Herzog, K.; et al. Candida albicans and early childhood caries: A systematic review and meta-analysis. Caries Res. 2018, 52, 102–112. [Google Scholar] [CrossRef] [PubMed]
  56. Bikash, C.R. Tal-Gan, Y. Structure activity relationship study of the XIP quorum sensing pheromone in Streptococcus mutans reveal inhibitors of the competence regulon. ACS Chem. Biol. 2020, 15, 2833–2841. [Google Scholar] [CrossRef]
  57. OmerOglou, E.; Karaca, B.; Kibar, H.; Haliscelik, O.; Kiran, F. The role of microbiota-derived postbiotic mediators on biofilm formation and quorum sensing-mediated virulence of Streptococcus mutans: A perspective on preventing dental caries. Microb. Pathog. 2022, 164, 105390. [Google Scholar] [CrossRef]
  58. Kaspar, J.R.; Walker, A.R. Expanding the vocabulary of peptide signals in Streptococcus mutans. Front. Cell Infect. Microbiol. 2019, 9, 194. [Google Scholar] [CrossRef]
  59. Suzuki, Y.; Nagasawa, R.; Senpuku, H. Inhibiting effects of fructanase on competence-stimulating peptide-dependent quorum sensing system in Streptococcus mutans. J. Infect. Chemother. 2017, 23, 634–641. [Google Scholar] [CrossRef]
  60. Okinaga, T.; Niu, G.; Xie, Z.; Qi, F.; Merritt, J. The hdrRM operon of Streptococcus mutans encodes a novel regulatory system for coordinated competence development and bacteriocin production. J. Bacteriol. 2010, 192, 1844–1852. [Google Scholar] [CrossRef]
  61. Underhill, S.A.M.; Shields, R.C.; Kaspar, J.R.; Haider, M.; Burne, R.A.; Hagen, S.J. Intracellular signaling by the comRS system in Streptococcus mutans genetic competence. Msphere 2018, 3, e00444-18. [Google Scholar] [CrossRef]
  62. Xie, Z.; Okinaga, T.; Niu, G.; Qi, F.; Merritt, J. Identification of a novel bacteriocin regulatory system in Streptococcus mutans. Mol. Microbiol. 2010, 78, 1431–1447. [Google Scholar] [CrossRef]
  63. Shanker, E.; Federle, M.J. Quorum sensing regulation of competence and bacteriocins in Streptococcus pneumoniae and mutans. Genes 2017, 8, 15. [Google Scholar] [CrossRef] [PubMed]
  64. Rued, B.E.; Covington, B.C.; Bushin, L.B.; Szewczyk, G.; Laczkovich, I.; Seyedsayamdost, M.R.; Federle, M.J. Quorum sensing in streptococcus mutans regulates production of tryglysin, a novel ras-ripp antimicrobial compound. MBio 2021, 12, e02688-20. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, X.; Li, X.; Ling, J. Streptococcus gordonii LuxS/autoinducer-2 quorum-sensing system modulates the dual-species biofilm formation with Streptococcus mutans. J. Basic Microbiol. 2017, 57, 605–616. [Google Scholar] [CrossRef]
  66. Yuan, K.; Hou, L.; Jin, Q.; Niu, C.; Mao, M.; Wang, R.; Huang, Z. Comparative transcriptomics analysis of Streptococcus mutans with disruption of LuxS/AI-2 quorum sensing and recovery of methyl cycle. Arch. Oral Biol. 2021, 127, 105137. [Google Scholar] [CrossRef] [PubMed]
  67. Kaspar, J.R.; Lee, K.; Richard, B.; Walker, A.R.; Burne, R.A. Direct interactions with commensal streptococci modify intercellular communication behaviors of Streptococcus mutans. ISME J. 2021, 15, 473–488. [Google Scholar] [CrossRef]
  68. Muadcheingka, T.; Tantivitayakul, P. Distribution of Candida albicans and non-albicans Candida species in oral candidiasis patients: Correlation between cell surface hydrophobicity and biofilm forming activities. Arch. Oral Biol. 2015, 60, 894–901. [Google Scholar] [CrossRef]
  69. Pereira, D.F.A.; Seneviratne, C.J.; Koga-Ito, C.Y.; Samaranayake, L.P. Is the oral fungal pathogen Candida albicans a cariogen? Oral Dis. 2018, 24, 518–526. [Google Scholar] [CrossRef]
  70. Lozano Moraga, C.P.; Rodríguez Martínez, G.A.; Lefimil Puente, C.A.; Morales Bozo, I.C.; Urzúa Orellana, B.R. Prevalence of Candida albicans and carriage of Candida non-albicans in the saliva of preschool children, according to their caries status. Acta Odontol. Scand. 2017, 75, 30–35. [Google Scholar] [CrossRef]
  71. Hornby, J.M.; Jensen, E.C.; Lisec, A.D.; Tasto, J.J.; Jahnke, B.; Shoemaker, R.; Dussault, P.; Nickerson, K.W. Quorum sensing in the dimorphic fungus Candida albicans is mediated by farnesol. Appl. Environ. Microbiol. 2001, 67, 2982–2992. [Google Scholar] [CrossRef]
  72. Petersen, F.C.; Scheie, A.A. Genetic transformation in Streptococcus mutans requires a peptide secretion-like apparatus. Oral Microbiol. Immunol. 2000, 15, 329–334. [Google Scholar] [CrossRef]
  73. Li, Y.H.; Lau, P.C.; Lee, J.H.; Ellen, R.P.; Cvitkovitch, D.G. Natural genetic transformation of Streptococcus mutans growing in biofilms. J. Bacteriol. 2001, 183, 897–908. [Google Scholar] [CrossRef] [PubMed]
  74. Biswas, S.; Cao, L.; Kim, A.; Biswas, I. SepM, a streptococcal protease involved in quorum sensing, displays strict substrate specificity. J. Bacteriol. 2016, 198, 436–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Li, Y.H.; Tang, N.; Aspiras, M.B.; Lau, P.C.; Lee, J.H.; Ellen, R.P.; Cvitkovitch, D.G. A quorum-sensing signaling system essential for genetic competence in Streptococcus mutans is involved in biofilm formation. J. Bacteriol. 2002, 184, 2699–2708. [Google Scholar] [CrossRef] [PubMed]
  76. Ishii, S.; Fukui, K.; Yokoshima, S.; Kumagai, K.; Beniyama, Y.; Kodama, T.; Fukuyama, T.; Okabe, T.; Nagano, T.; Kojima, H.; et al. High.-throughput screening of small molecule inhibitors of the Streptococcus quorum-sensing signal pathway. Sci. Rep. 2017, 7, 4029. [Google Scholar] [CrossRef]
  77. Kaspar, J.; Underhill, S.A.; Shields, R.C.; Reyes, A.; Rosenzweig, S.; Hagen, S.J.; Burne, R.A. Intercellular communication via the comX-inducing peptide (XIP) of Streptococcus mutans. J. Bacteriol. 2017, 199, e00404-17. [Google Scholar] [CrossRef]
  78. Mashburn-Warren, L.; Morrison, D.A.; Federle, M.J. A novel double-tryptophan peptide pheromone controls competence in Streptococcus spp. via an Rgg regulator. Mol. Microbiol. 2010, 78, 589–606. [Google Scholar] [CrossRef]
  79. Talagas, A.; Fontaine, L.; Ledesma-García, L.; Mignolet, J.; Li de la Sierra-Gallay, I.; Lazar, N.; Aumont-Nicaise, M.; Federle, M.J.; Prehna, G.; Hols, P.; et al. Structural insights into streptococcal competence regulation by the cell-to-cell communication system ComRS. PLoS Pathog. 2016, 12, e1005980. [Google Scholar] [CrossRef]
  80. Ledesma-García, L.; Ensinck, I.; Dereinne, D.; Viela, F.; Mignolet, J.; Dufrêne, Y.F.; Soumillion, P.; Nessler, S.; Hols, P. Coevolution of the bacterial pheromone ComS and sensor ComR fine-tunes natural transformation in streptococci. J. Biol. Chem. 2021, 297, 101346. [Google Scholar] [CrossRef]
  81. Zhang, Z.; Liu, Y.; Lu, M.; Lyu, X.; Gong, T.; Tang, B.; Wang, L.; Zeng, J.; Li, Y. Rhodiola rosea extract inhibits the biofilm formation and the expression of virulence genes of cariogenic oral pathogen Streptococcus mutans. Arch. Oral Biol. 2020, 116, 104762. [Google Scholar] [CrossRef]
  82. Balasubramanian, A.; Vasudevan, S.; Shanmugam, K.; Lévesque, C.; Solomon, A.; Neelakantan, P. Combinatorial effects of trans-cinnamaldehyde with fluoride and chlorhexidine on Streptococcus mutans. J. Appl. Microbiol. 2021, 130, 382–393. [Google Scholar] [CrossRef]
  83. 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] [PubMed]
  84. Eguchi, Y.; Kubo, N.; Matsunaga, H.; Igarashi, M.; Utsumi, R. Development of an antivirulence drug against Streptococcus mutans: Repression of biofilm formation, acid tolerance, and competence by a histidine kinase inhibitor, walkmycin C. Antimicrob. Agents Chemother. 2011, 55, 1475–1484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Ogawa, A.; Furukawa, S.; Fujita, S.; Mitobe, J.; Kawarai, T.; Narisawa, N.; Sekizuka, T.; Kuroda, M.; Ochiai, K.; Ogihara, H.; et al. Inhibition of Streptococcus mutans biofilm formation by Streptococcus salivarius FruA. Appl. Env. Microbiol. 2011, 77, 1572–1580. [Google Scholar] [CrossRef] [PubMed]
  86. Wasfi, R.; Abd El-Rahman, O.A.; Zafer, M.M.; Ashour, H.M. Probiotic Lactobacillus sp. inhibit growth, biofilm formation and gene expression of caries-inducing Streptococcus mutans. J. Cell Mol. Med. 2018, 22, 1972–1983. [Google Scholar] [CrossRef] [PubMed]
  87. Pourhajibagher, M.; Alaeddini, M.; Etemad-Moghadam, S.; Rahimi Esboei, B.; Bahrami, R.; Mousavi, M.; Bahador, A. Quorum quenching of Streptococcus mutans via the nano-quercetin-based antimicrobial photodynamic therapy as a potential target for cariogenic biofilm. BMC Microbiol. 2022, 22, 1–18. [Google Scholar] [CrossRef]
  88. Davis-Hanna, A.; Piispanen, A.E.; Stateva, L.I.; Hogan, D.A. Farnesol and dodecanol effects on the Candida albicans Ras1-cAMP signalling pathway and the regulation of morphogenesis. Mol. Microbiol. 2008, 67, 47–62. [Google Scholar] [CrossRef]
  89. Sebaa, S.; Boucherit-Otmani, Z.; Courtois, P. Effects of tyrosol and farnesol on Candida albicans biofilm. Mol. Med. Rep. 2019, 19, 3201–3209. [Google Scholar] [CrossRef]
  90. Kim, D.; Sengupta, A.; Niepa, T.H.; Lee, B.H.; Weljie, A.; Freitas-Blanco, V.S.; Murata, R.M.; Stebe, K.J.; Lee, D.; Koo, H. Candida albicans stimulates Streptococcus mutans microcolony development via cross-kingdom biofilm-derived metabolites. Sci. Rep. 2017, 7, 41332. [Google Scholar] [CrossRef]
  91. Bowen, W.H.; Koo, H. Biology of Streptococcus mutans-derived glucosyltransferases: Role in extracellular matrix formation of cariogenic biofilms. Caries Res. 2011, 45, 69–86. [Google Scholar] [CrossRef]
  92. Sztajer, H.; Szafranski, S.P.; Tomasch, J.; Reck, M.; Nimtz, M.; Rohde, M.; Wagner-Döbler, I. Cross-feeding and interkingdom communication in dual-species biofilms of Streptococcus mutans and Candida albicans. ISME J. 2014, 8, 2256–2271. [Google Scholar] [CrossRef]
  93. He, J.; Kim, D.; Zhou, X.; Ahn, S.J.; Burne, R.A.; Richards, V.P.; Koo, H. RNA-Seq reveals enhanced sugar metabolism in Streptococcus mutans co-cultured with Candida albicans within mixed-species biofilms. Front. Microbiol. 2017, 8, 1036. [Google Scholar] [CrossRef] [PubMed]
  94. Lobo, C.I.V.; Lopes, A.C.U.d.A.; Klein, M.I. Compounds with distinct targets present diverse antimicrobial and antibiofilm efficacy against Candida albicans and Streptococcus mutans, and combinations of compounds potentiate their effect. J. Fungi. 2021, 7, 340. [Google Scholar] [CrossRef] [PubMed]
  95. Priya, A.; Selvaraj, A.; Divya, D.; Karthik Raja, R.; Pandian, S.K. In vitro and in vivo anti-infective potential of thymol against early childhood caries causing dual species Candida albicans and Streptococcus mutans. Front. Pharmacol. 2021, 12, 760768. [Google Scholar] [CrossRef]
  96. Alemán-González-Duhart, D.; Álvarez-Almazán, S.; Valdes, M.; Tamay-Cach, F.; Mendieta-Wejebe, J.E. In vivo and ex vivo evaluation of 1,3-thiazolidine-2,4-dione derivatives as euglycemic agents. PPAR Res. 2021, 2021, 5100531. [Google Scholar] [CrossRef]
  97. Feldman, M.; Shenderovich, J.; Lavy, E.; Friedman, M.; Steinberg, D. A sustained-release membrane of thiazolidinedione-8: Effect on formation of a candida/bacteria mixed biofilm on hydroxyapatite in a continuous flow model. Biomed. Res. Int. 2017, 2017, 3510124. [Google Scholar] [CrossRef]
  98. Yost, S.; Duran-Pinedo, A.E.; Teles, R.; Krishnan, K.; Frias-Lopez, J. Functional signatures of oral dysbiosis during periodontitis progression revealed by microbial metatranscriptome analysis. Genome Med. 2015, 7, 27. [Google Scholar] [CrossRef] [PubMed]
  99. Frias, J.; Olle, E.; Alsina, M. Periodontal pathogens produce quorum sensing signal molecules. Infect. Immun. 2001, 69, 3431–3434. [Google Scholar] [CrossRef] [PubMed]
  100. Hajishengallis, G.; Darveau, R.P.; Curtis, M.A. The keystone-pathogen hypothesis. Nat. Rev. Microbiol. 2012, 10, 717–725. [Google Scholar] [CrossRef]
  101. Lamont, R.J.; Koo, H.; Hajishengallis, G. The oral microbiota: Dynamic communities and host interactions. Nat. Rev. Microbiol. 2018, 16, 745–759. [Google Scholar] [CrossRef]
  102. Socransky, S.S.; Haffajee, A.D.; Cugini, M.A.; Smith, C.; Kent, R.L., Jr. Microbial complexes in subgingival plaque. J. Clin. Periodontol. 1998, 25, 134–144. [Google Scholar] [CrossRef]
  103. Thurnheer, T.; Karygianni, L.; Flury, M.; Belibasakis, G.N. Fusobacterium species and subspecies differentially affect the composition and architecture of supra-and subgingival biofilms models. Front. Microbiol. 2019, 10, 1716. [Google Scholar] [CrossRef] [PubMed]
  104. Muras, A.; Mayer, C.; Otero-Casal, P.; Exterkate, R.A.M.; Brandt, B.W.; Crielaard, W.; Otero, A.; Krom, B.P. Short-chain N-acylhomoserine lactone quorum-sensing molecules promote periodontal pathogens in in vitro oral biofilms. Appl. Environ. Microbiol 2020, 86, e01941-19. [Google Scholar] [CrossRef] [PubMed]
  105. Jang, Y.J.; Choi, Y.J.; Lee, S.H.; Jun, H.K.; Choi, B.K. Autoinducer 2 of Fusobacterium nucleatum as a target molecule to inhibit biofilm formation of periodontopathogens. Arch. Oral Biol. 2013, 58, 17–27. [Google Scholar] [CrossRef] [PubMed]
  106. Jang, Y.J.; Sim, J.; Jun, H.K.; Choi, B.K. Differential effect of autoinducer 2 of Fusobacterium nucleatum on oral streptococci. Arch. Oral Biol. 2013, 58, 1594–1602. [Google Scholar] [CrossRef]
  107. Zhang, Y.; Shi, W.; Song, Y.; Wang, J. Metatranscriptomic analysis of an in vitro biofilm model reveals strain-specific interactions among multiple bacterial species. J. Oral Microbiol. 2019, 11, 1599670. [Google Scholar] [CrossRef]
  108. Rupani, D.; Izano, E.A.; Schreiner, H.C.; Fine, D.H.; Kaplan, J.B. Aggregatibacter actinomycetemcomitans serotype f O-polysaccharide mediates coaggregation with Fusobacterium nucleatum. Oral Microbiol. Immunol. 2008, 23, 127–130. [Google Scholar] [CrossRef]
  109. Azakami, H.; Teramura, I.; Matsunaga, T.; Akimichi, H.; Noiri, Y.; Ebisu, S.; Kato, A. Characterization of autoinducer 2 signal in Eikenella corrodens and its role in biofilm formation. J. Biosci. Bioeng. 2006, 102, 110–117. [Google Scholar] [CrossRef]
  110. Szafrański, S.P.; Deng, Z.L.; Tomasch, J.; Jarek, M.; Bhuju, S.; Rohde, M.; Sztajer, H.; Wagner-Döbler, I. Quorum sensing of Streptococcus mutans is activated by Aggregatibacter actinomycetemcomitans and by the periodontal microbiome. BMC Genom. 2017, 18, 238. [Google Scholar] [CrossRef]
  111. Novak, E.A.; Shao, H.; Daep, C.A.; Demuth, D.R. Autoinducer-2 and QseC control biofilm formation and in vivo virulence of Aggregatibacter actinomycetemcomitans. Infect. Immun. 2010, 78, 2919–2926. [Google Scholar] [CrossRef]
  112. Ponnusamy, K.; Paul, D.; Sam Kim, Y.; Kweon, J.H. 2(5H)-Furanone: A prospective strategy for biofouling-control in membrane biofilm bacteria by quorum sensing inhibition. Braz. J. Microbiol. 2010, 41, 227–234. [Google Scholar] [CrossRef]
  113. Rasmussen, T.B.; Manefield, M.; Andersen, J.B.; Eberl, L.; Anthoni, U.; Christophersen, C.; Steinberg, P.; Kjelleberg, S.; Givskov, M. How Delisea pulchra furanones affect quorum sensing and swarming motility in Serratia liquefaciens MG1. Microbiology 2000, 146 Pt 12, 3237–3244. [Google Scholar] [CrossRef]
  114. Park, J.S.; Ryu, E.J.; Li, L.; Choi, B.K.; Kim, B.M. New bicyclic brominated furanones as potent autoinducer-2 quorum-sensing inhibitors against bacterial biofilm formation. Eur. J. Med. Chem. 2017, 137, 76–87. [Google Scholar] [CrossRef]
  115. Zhang, L.; Wang, S.; Zhou, X.; Xui, Y. Quorum sensing inhibitor brominated furanone affects Porphyromonas gingivalis biofilm formation. West. China J. Stomatol. 2011, 29, 469–472. [Google Scholar]
  116. Manny, A.J.; Kjelleberg, S.; Kumar, N.; de Nys, R.; Read, R.W.; Steinberg, P. Reinvestigation of the sulfuric acid-catalysed cyclisation of brominated 2-alkyllevulinic acids to 3-alkyl-5-methylene-2 (5H)-furanones. Tetrahedron 1997, 53, 15813–15826. [Google Scholar] [CrossRef]
  117. Cho, Y.J.; Song, H.Y.; Ben Amara, H.; Choi, B.K.; Eunju, R.; Cho, Y.A.; Seol, Y.; Lee, Y.; Ku, Y.; Rhyu, I.C.; et al. In vivo inhibition of Porphyromonas gingivalis growth and prevention of periodontitis with quorum-sensing inhibitors. J. Periodontol. 2016, 87, 1075–1082. [Google Scholar] [CrossRef] [PubMed]
  118. Ryu, E.J.; Sim, J.; Sim, J.; Lee, J.; Choi, B.K. D-Galactose as an autoinducer 2 inhibitor to control the biofilm formation of periodontopathogens. J. Microbiol. 2016, 54, 632–637. [Google Scholar] [CrossRef]
  119. Fournier-Larente, J.; Morin, M.P.; Grenier, D. Green tea catechins potentiate the effect of antibiotics and modulate adherence and gene expression in Porphyromonas gingivalis. Arch. Oral Biol. 2016, 65, 35–43. [Google Scholar] [CrossRef]
  120. Asahi, Y.; Noiri, Y.; Igarashi, J.; Asai, H.; Suga, H.; Ebisu, S. Effects of N-acyl homoserine lactone analogues on Porphyromonas gingivalis biofilm formation. J. Periodontal. Res. 2010, 45, 255–261. [Google Scholar] [CrossRef]
  121. Elani, H.W.; Starr, J.R.; Da Silva, J.D.; Gallucci, G.O. Trends in dental implant use in the U.S., 1999–2016, and projections to 2026. J. Dent. Res. 2018, 97, 1424–1430. [Google Scholar] [CrossRef]
  122. Ramachandra, S.S.; Patil, M.; Mehta, D.S. Implants placed into extraction sockets: A literature review. Dent. Implant. Update 2009, 20, 1–8. [Google Scholar]
  123. Ramachandra, S.S.; Rana, R.; Reetika, S.; Jithendra, K.D. Options to avoid the second surgical site: A review of literature. Cell Tissue Bank 2014, 15, 297–305. [Google Scholar] [CrossRef] [PubMed]
  124. Heitz-Mayfield, L.J.A.; Salvi, G.E. Peri-implant mucositis. J. Clin. Periodontol. 2018, 45 (Suppl. 20), S237–S245. [Google Scholar] [CrossRef] [PubMed]
  125. Schwarz, F.; Derks, J.; Monje, A.; Wang, H.L. Peri-implantitis. J. Clin. Periodontol. 2018, 45 (Suppl. 20), S246–S266. [Google Scholar] [CrossRef] [PubMed]
  126. Renvert, S.; Persson, G.R.; Pirih, F.Q.; Camargo, P.M. Peri-implant health, peri-implant mucositis, and peri-implantitis: Case definitions and diagnostic considerations. J. Clin. Periodontol. 2018, 45 (Suppl. 20), S278–S285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Nicolae, V.D.; Chiscop, I.; Cioranu, V.S.I.; Mârțu, M.-A.; Luchian, A.I.; Mârțu, S.; Solomon, S.M. The use of photoactivated blue-O toluidine for periimplantitis treatment in patients with periodontal disease. Rev. Chim. (Buchar.) 2016, 66, 2121–2123. [Google Scholar]
  128. Kang, M.; Kim, S.; Kim, H.; Song, Y.; Jung, D.; Kang, S.; Seo, J.H.; Nam, S.; Lee, Y. Calcium-binding polymer-coated poly(lactide- co-glycolide) microparticles for sustained release of quorum sensing inhibitors to prevent biofilm formation on hydroxyapatite surfaces. ACS Appl. Mater. Interfaces 2019, 11, 7686–7694. [Google Scholar] [CrossRef]
  129. An, S.J.; Namkung, J.U.; Ha, K.W.; Jun, H.K.; Kim, H.Y.; Choi, B.K. Inhibitory effect of D-arabinose on oral bacteria biofilm formation on titanium discs. Anaerobe 2022, 102533. [Google Scholar] [CrossRef]
  130. Rajkumar, D.; Rubini, D.; Sudharsan, M.; Suresh, D.; Nithyanand, P. Novel thiazolinyl-picolinamide based palladium(II) complex-impregnated urinary catheters quench the virulence and disintegrate the biofilms of uropathogens. Biofouling 2020, 36, 351–367. [Google Scholar] [CrossRef]
  131. Jothipandiyan, S.; Suresh, D.; Sankaran, S.V.; Thamotharan, S.; Shanmugasundaram, K.; Vincent, P.; Sekaran, S.; Gowrishankar, S.; Pandian, S.K.; Paramasivam, N. Heteroleptic pincer palladium(II) complex coated orthopedic implants impede the AbaI/AbaR quorum sensing system and biofilm development by Acinetobacter baumannii. Biofouling 2022, 38, 55–70. [Google Scholar] [CrossRef]
  132. Ganesan, S.M.; Dabdoub, S.M.; Nagaraja, H.N.; Mariotti, A.J.; Ludden, C.W.; Kumar, P.S. Biome-microbiome interactions in peri-implantitis: A pilot investigation. J. Periodontol. 2022, 93, 814–823. [Google Scholar] [CrossRef]
  133. Kumar, P.S.; Mason, M.R.; Brooker, M.R.; O’Brien, K. Pyrosequencing reveals unique microbial signatures associated with healthy and failing dental implants. J. Clin. Periodontol. 2012, 39, 425–433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. QS model in Gram-negative bacteria. Feedback loop involving the catalysis of acyl-homoserine lactone (AHL) by LuxI from S-adenosylmethionine (SAM) and fatty acid synthesis intermediaries. Upon re-entry into the cell, AHL combines with LuxR to initiate the expression of virulence factors, LuxI and LuxR.
Figure 1. QS model in Gram-negative bacteria. Feedback loop involving the catalysis of acyl-homoserine lactone (AHL) by LuxI from S-adenosylmethionine (SAM) and fatty acid synthesis intermediaries. Upon re-entry into the cell, AHL combines with LuxR to initiate the expression of virulence factors, LuxI and LuxR.
Microorganisms 10 01783 g001
Figure 2. QS in Gram-positive bacteria. AIP (autoinducing peptide) is exported extracellularly, where at increased cell density, enough accumulates to be sensed by the histidine kinase receptor to initiate the transcription of AIP precursors, virulence factors, an export channel, and the regulator and receptor proteins.
Figure 2. QS in Gram-positive bacteria. AIP (autoinducing peptide) is exported extracellularly, where at increased cell density, enough accumulates to be sensed by the histidine kinase receptor to initiate the transcription of AIP precursors, virulence factors, an export channel, and the regulator and receptor proteins.
Microorganisms 10 01783 g002
Figure 4. Implant abutment surfaces are common areas where bacteria accumulate, resulting in the induction of quorum sensing through the release of autoinducer molecules. Robust virulent biofilms are formed which may cause peri-implantitis. Implant abutment surfaces can be coated with QS inhibitors resulting in quorum quenching. This inhibits biofilm formation.
Figure 4. Implant abutment surfaces are common areas where bacteria accumulate, resulting in the induction of quorum sensing through the release of autoinducer molecules. Robust virulent biofilms are formed which may cause peri-implantitis. Implant abutment surfaces can be coated with QS inhibitors resulting in quorum quenching. This inhibits biofilm formation.
Microorganisms 10 01783 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wright, P.P.; Ramachandra, S.S. Quorum Sensing and Quorum Quenching with a Focus on Cariogenic and Periodontopathic Oral Biofilms. Microorganisms 2022, 10, 1783. https://doi.org/10.3390/microorganisms10091783

AMA Style

Wright PP, Ramachandra SS. Quorum Sensing and Quorum Quenching with a Focus on Cariogenic and Periodontopathic Oral Biofilms. Microorganisms. 2022; 10(9):1783. https://doi.org/10.3390/microorganisms10091783

Chicago/Turabian Style

Wright, Patricia P., and Srinivas Sulugodu Ramachandra. 2022. "Quorum Sensing and Quorum Quenching with a Focus on Cariogenic and Periodontopathic Oral Biofilms" Microorganisms 10, no. 9: 1783. https://doi.org/10.3390/microorganisms10091783

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